[3]Symplectic Geometric Algorithms for Hamiltonian Systems
[3]Symplectic Geometric Algorithms for Hamiltonian Systems
Mengzhao Qin
Symplectic Geometric
Algorithms for
Hamiltonian Systems
With 62 Figures
ISBN 978-7-5341-3595-8
Zhejiang Publishing United Group, Zhejiang Science and Technology Publishing House,
Hangzhou
¤ Zhejiang Publishing United Group, Zhejiang Science and Technology Publishing House,
Hangzhou and Springer-Verlag Berlin Heidelberg 2010
This work is subject to copyright. All rights are reserved, whether the whole or part of the material
is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation,
broadcasting, reproduction on microfilm or in any other way, and storage in data banks.
Duplication of this publication or parts thereof is permitted only under the provisions of the
German Copyright Law of September 9, 1965, in its current version, and permission for use must
always be obtained from Springer. Violations are liable to prosecution under the German Copyright
Law.
The use of general descriptive names, registered names, trademarks, etc. in this publication does not
imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
Peter Lax
Cited from a paper entitled “How to compute property Newton’s equation of motion”
Prize certificate
Kang Feng (1920–1993), Member of the Chinese Academy of Sciences, Professor and
Honorary Director of the Computing Center of the Chinese Academy of Sciences,
famous applied mathematician, founder and pioneer of computational mathematics
and scientific computing in China.
It has been 16 years since my brother Kang Feng passed away. His scientific
achievements have been recognized more and more clearly over time, and his contri-
butions to various fields have become increasingly outstanding. In the spring of 1997,
Professor Shing-Tung Yau, a winner of the Fields Medal and a foreign member of the
Chinese Academy of Sciences, mentioned in a presentation at Tsinghua University,
entitled “The development of mathematics in China in my view”, that “there are three
main reasons for Chinese modern mathematics to go beyond or hand in hand with
the West. Of course, I am not saying that there are no other works, but I mainly talk
about the mathematics that is well known historically: Professor Shiingshen Chern’s
work on characteristic class, Luogeng Hua’s work on the theory of functions of several
complex variables, and Kang Feng’s work on finite elements.” This high evaluation of
Kang Feng as a mathematician (not just a computational mathematician) sounds so
refreshing that many people talked about it and strongly agreed with it. At the end
of 1997, the Chinese National Natural Science Foundation presented Kang Feng et
al. with the first class prize for his other work on a symplectic algorithm for Hamil-
tonian systems, which is a further recognition of his scientific achievements (see the
certificate on the previous page). As his brother, I am very pleased.
Achieving a major scientific breakthrough is a rare event. It requires vision, ability
and opportunity, all of which are indispensable. Kang Feng has achieved two major
scientific breakthroughs in his life, both of which are very valuable and worthy of
mention. Firstly, from 1964 to 1965, he proposed independently the finite element
method and laid the foundation for the mathematical theory. Secondly, in 1984, he
proposed a symplectic algorithm for Hamiltonian systems. At present, scientific inno-
vation has become the focus of discussion. Kang Feng’s two scientific breakthroughs
may be treated as case studies in scientific innovation. It is worth emphasizing that
these breakthroughs were achieved in China by Chinese scientists. Careful study of
these has yet to be carried out by experts. Here I just describe some of my personal
feelings.
It should be noted that these breakthroughs resulted not only from the profound
mathematical knowledge of Kang Feng, but also from his expertise in classical physics
and engineering technology that were closely related to the projects. Scientific break-
throughs are often cross-disciplinary. In addition, there is often a long period of time
before a breakthrough is made-not unlike a long time it takes for a baby to be born,
which requires the accumulation of results in small steps.
x Foreword
The opportunity for inventing the finite element method came from a national re-
search project, a computational problem in the design of the Liu Jia Xia dam. For
such a concrete problem, Kang Feng found a basis for solving of the problem using
his sharp insight. In his view, a discrete computing method for a mathematical and
physical problem is usually carried out in four steps. Firstly, one needs to know and
define the physical mechanism. Secondly, one writes the appropriate differential equa-
tions accordingly. In the third step, design a discrete model. Finally, one develops the
numerical algorithm. However, due to the complexity of the geometry and physical
conditions, conventional methods cannot always be effective. Nonetheless, starting
from the physical law of conservation or variational principle of the matter, we can
directly relate to the appropriate discrete model. Combining the variational principle
with the spline approximation leads to the finite element method, which has a wide
range of adaptability and is particularly suited to deal with the complex geometry of
the physical conditions of computational engineering problems. In 1965, Kang Feng
published his paper entitled “Difference schemes based on the variational principle”,
which solved the basic theoretical issues of the finite element method, such as conver-
gence, error estimation, and stability. It laid the mathematical foundation for the finite
element method. This paper is the main evidence for recognition by the international
academic community of our independent development of the finite element method.
After the Chinese Cultural Revolution, he continued his research in finite element
and related areas. During this period, he made several great achievements. I remem-
ber that he talked with me about other issues, such as Thom’s catastrophe theory,
Prigogine’s theory of dissipative structures, solitons in water waves, the Radon trans-
form, and so on. These problems are related to physics and engineering technology.
Clearly he was exploring for new areas and seeking a breakthrough. In the 1970s,
Arnold’s “Mathematical Method of Classical Mechanics” came out. It described the
symplectic structure for Hamiltonian equations, which proved to be a great inspira-
tion to him and led to a breakthrough. Through his long-term experience in mathe-
matical computation, he fully realized that different mathematical expressions for the
same physical law, which are physically equivalent, can perform different functions
in scientific computing (his students later called this the “Feng’s major theorem”).
In this way, for classical mechanics, Newton’s equations, Lagrangian equations and
Hamiltonian equations will show a different pattern of calculations after discretiza-
tion. Because the Hamiltonian formulation has a symplectic structure, he was keenly
aware that, if the algorithm can maintain the geometric symmetry of symplecticity, it
will be possible to avoid the flaw of artificial dissipation of this type of algorithm and
design a high-fidelity algorithm. Thus, he opened up a broad way for the computa-
tional method of the Hamiltonian system. He called this way the “Hamiltonian way”.
This computational method has been used in the calculation of the orbit in celestial
mechanics, in calculations for the particle path in accelerator, as well as in molecular
dynamics. Later, the scope of its application was expanded. For example, it has also
been widely used in studies of the atmosphere and earth sciences and elsewhere. It
Foreword xi
has been effectively applied in solving the GPS observation operator, indicating that
Global Positioning System data can be dealt with in a timely manner. This algorithm
is 400 times more efficient than the traditional method. In addition, a symplectic al-
gorithm has been successfully used in the oil and gas exploration fields. Under the
influence of Kang Feng, international research on symplectic algorithm has become
popular and flourishing, nearly 300 papers have been published in this field to date.
Kang Feng’s research work on the symplectic algorithm has been well-known and
recognized internationally for its unique, innovative, systematic and widespread prop-
erties, for its theoretical integrity and fruitful results.
J. Lions, the former President of the International Mathematics Union, spoke at
a workshop when celebrating his 60th birthday: “This is another major innovation
made by Kang Feng, independent of the West, after the finite element method.” In
1993 one of the world’s leading mathematicians, P.D. Lax, a member of the Ameri-
can Academy of Sciences, wrote a memorial article dedicated to Kang Feng in SIAM
News, stating that “In the late 1980s, Kang Feng proposed and developed so-called
symplectic algorithms for solving evolution equations . . .. Such methods, over a long
period, are much superior to standard methods.” E. J. Marsden, an internationlly well-
known applied mathematician, visited the computing institute in the late 1980s and
had a long conversation with Kang Feng. Soon after the death of Kang Feng, he pro-
posed the multi-symplectic algorithm and extended the characteristics of stability of
the symplectic algorithm for long time calculation of Hamiltonian systems with infi-
nite dimensions.
On the occasion of the commemoration of the 16th anniversary of Kang Feng’s
death and the 89th anniversary of his birth, I think it is especially worthwhile to praise
and promote what was embodied in the lifetime’s work of Kang Feng — “ indepen-
dence in spirit, freedom in thinking”. 1 Now everyone is talking about scientific inno-
vation, which needs a talented person to accomplish. What type of person is needed
most? A person who is just a parrot or who has an “independent spirit, freely think-
ing”? The conclusion is self-evident. Scientific innovation requires strong academic
atmosphere. Is it determined by only one person or by all of the team members? This
is also self-evident. From Kang Feng’s scientific career, we can easily find that the key
to the problem of scientific innovation is “independence in spirit, freedom in thinking”,
and that needs to be allowed to develop and expand.
Kang Feng had planned to write a monograph about a symplectic algorithm for
Hamiltonian systems. He had accumulated some manuscripts, but failed to complete
it because he died too early due to sickness. Fortunately, his students and Professor
Mengzhao Qin (see the photo on the previous page), one of the early collaborators,
spent 15 years and finally completed this book based on Kang Feng’s plan, realizing
his wish. It is not only an authoritative exposition of this research field, but also an
1
Yinke Chen engraved on a stele in 1929 in memory of Guowei Wang in campus of Tsinghua
University.
xii Foreword
Duan, Feng
Member of Chinese
Academy of Sciences
Nanjing University
Nanjing
September 20, 2009
Preface
It has been 16 years since Kang Feng passed away. It is our honor to publish the En-
glish version of Symplectic Algorithm for Hamiltonian Systems, so that more readers
can see the history of the development of symplectic algorithms. In particular, after
the death of Kang Feng, the development of symplectic algorithms became more so-
phisticated and there have been a series of monographs published in this area, e.g.,
Sanz-Serna & M.P. Calvo’s Numerical Hamiltonian Problems published in 1994 by
Chapman and Hall Publishing House; E. Hairer, C. Lubich and G. Wanner’s Geo-
metrical Numerical Integration published in 2001 by Springer Verlag; B. Leimkuhler
and S. Reich’s Simulating Hamiltonian Dynamics published in 2004 by Cambridge
University Press. The symplectic algorithm has been developed from ordinary dif-
ferential equations to partial differential equations, from a symplectic structure to a
multi-symplectic structure. This is largely due to the promotion of this work by J.
Marsden of the USA and T. Bridge and others in Britain. Starting with a symplectic
structure, J. Marsden first developed the Lagrange symplectic structure, and then to
the multi-symplectic structure. He finally proposed a symplectic structure that meets
the requirement of the Lagrangian form from the variational principle by giving up
the boundary conditions. On the other hand, T. Bridge and others used the multi-
symplectic structure to derive directly the multi-symplectic Hamilton equations, and
then constructed the difference schemes that preserve the symplectic structure in both
time and space. Both methods can be regarded as equivalent in the algorithmic sense.
Now, in this monograph, most of the content refers only to ordinary differential
equations. Kang Feng and his algorithms research group working on the symplectic
algorithm did some foundation work. In particular, I would like to point out three nega-
tive theorems: “ non-existence of energy preserving scheme”, “ non-existence of mul-
tistep linear symplectic scheme”, and “ non-existence of volume-preserving scheme
form rational fraction expression”. In addition, generating function theory is not only
rich in analytical mechanics and Hamilton–Jacobi equations. At the same time, the
construction of symplectic schemes provides a tool for any order accuracy difference
scheme. The formal power series proposed by Kang Feng had a profound impact on
the later developed “ backward error series” work ,“ modified equation” and “ modified
integrator”.
The symplectic algorithm developed very quickly, soon to be extended to the ge-
ometric method. The structure preserving algorithm (not only preserving the geomet-
rical structure, but also the physical structure, etc.) preserves the algebraic structure
to present the Lie group algorithm, and preserves the differential complex algorithm.
Many other prominent people have contributed to the symplectic method in addition
to those mentioned above. There are various methods related to structure preserving
algorithms and for important contributions the readers are referred to R. McLach-
lan & GRW Quispel “ Geometric integration for ODEs” and T. Bridges & S. Reich
“ Numerical methods for Hamiltonian PDEs”.
The book describes the symplectic geometric algorithms and theoretical basis for
a number of related algorithms. Most of the contents are a collection of lectures given
xiv Preface
by Kang Feng at Beijing University. Most of other sections are a collection of papers
which were written by group members.
Compared to the previous Chinese version, the present English one has been im-
proved in the following respects. First of all, to correct a number of errors and mis-
takes contained in the Chinese version. Besides, parts of Chapter 1 and Chapter 2
were removed, while some new content was added to Chapter 4, Chapter 7, Chapter
8, Chapter 9 and Chapter 10. More importantly, four new chapters — Chapter 13 to
Chapter 16 were added. Chapter 13 is devoted to the KAM theorem for the symplectic
algorithm. We invited Professor Zaijiu Shang , a former PhD student of Kang Feng
to compose this chapter. Chapter 14 is called Variational Integrator. This chapter re-
flects the work of the Nobel Prize winner Professor Zhengdao Li who proposed in
the 1980s to preserve the energy variational integrator, but had not explained at that
time that it had a Lagrange symplectic type, which satisfied the Lagrange symplectic
structure. Together with J. Marsden he proposed the variational integrator trail con-
nection, which leads from the variational integrator. Just like J. Marsden, he hoped
this can link up with the finite element method. Chapter 15 is about Birkhoffian Sys-
tems, describing a class of dissipative structures for Birkohoffian systems to preserve
the dissipation of the Birkhoff structure. Chapter 16 is devoted to Multisymplectic
and Variational Integrators, providing a summary of the widespread applications of
multisymplectic integrators in the infinitely dimensional Hamiltonian systems.
We would also like to thank every member of the Kang Feng’s research group
for symplectic algorithms: Huamo Wu, Daoliu Wang, Zaijiu Shang, Yifa Tang, Jialin
Hong, Wangyao Li, Min Chen, Shuanghu Wang, Pingfu Zhao, Jingbo Chen, Yushun
Wang, Yajuan Sun, Hongwei Li, Jianqiang Sun, Tingting Liu, Hongling Su, Yimin
Tian; and those who have been to the USA: Zhong Ge, Chunwang Li, Yuhua Wu,
Meiqing Zhang, Wenjie Zhu, Shengtai Li, Lixin Jiang, and Haibin Shu. They made
contributions to the symplectic algorithm over different periods of time.
The authors would also like to thank the National Natural Science Foundation, the
National Climbing Program projects, and the State’s Key Basic Research Projects for
their financial support. Finally, the authors would also like to thank the Mathematics
and Systems Science Research Institute of the Chinese Academy of Sciences, the
Computational Mathematics and Computational Science and Engineering Institute,
and the State Key Laboratory of Computational Science and Engineering for their
support.
The editors of this book have received help from E. Hairer, who provided a tem-
plate from Springer publishing house. I would also like to thank F. Holzwarth at
Springer publishing house and Linbo Zhang of our institute, and others who helped
me successfully publish this book.
For the English translation, I thank Dr. Shengtai Li for comprehensive proof-
reading and polishing, and the editing of Miss Yi Jin. For the English version of the
publication I would also like to thank the help of the Chinese Academy of Sciences
Institute of Mathematics. Because Kang Feng has passed away, it may not be possible
to provide a comprehensive representation of his academic thought, and the book will
inevitably contain some errors. I accept the responsibility for any errors and welcome
criticism and corrections.
Preface xv
We would also like to thank Springer Beijing Representation Office and Zhejiang
Science and Technology Publishing House, which made a great contribution to the
Chinese scientific cause through the publication of this manuscript. We are especially
grateful to thank Lisa Fan, W. Y. Zhou, L. L. Liu and X. M. Lu for carefully reading
and finding some misprints, wrong signs and other mistakes.
This book is supported by National Natural Science Foundation of China under
grant No.G10871099 ; supported by the Project of National 863 Plan of China (grant
No.2006AA09A102-08); and supported by the National Basic Research Program of
China (973 Program) (Grant No. 2007CB209603).
Mengzhao Qin
Institute of Computational
Mathematics and Scientific
Engineering Computing
Beijing
September 20, 2009
Contents
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
Symbol . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 669
Introduction
The main theme of modern scientific computing is the numerical solution of various
differential equations of mathematical physics bearing the names, such as Newton, Eu-
ler, Lagrange, Laplace, Navier–Stokes, Maxwell, Boltzmann, Einstein, Schrödinger,
Yang-Mills, etc. At the top of the list is the most celebrated Newton’s equation of mo-
tion. The historical, theoretical and practical importance of Newton’s equation hardly
needs any comment, so is the importance of the numerical solution of such equations.
On the other hand, starting from Euler, right down to the present computer age, a
great wealth of scientific literature on numerical methods for differential equations
has been accumulated, and a great variety of algorithms, software packages and even
expert systems has been developed. With the development of the modern mechanics,
physics, chemistry, and biology, it is undisputed that almost all physical processes,
whether they are classical, quantum, or relativistic, can be represented by an Hamilto-
nian system. Thus, it is important to solve the Hamiltonian system correctly.
latter transforms the second-order differential equations in physical space into a group
of the first-order canonical equations in phase space. Different representations for
the same physical law can lead to different computational techniques in solving
the same problem, which can produce different numerical results. Thus making a
wise and reasonable choice among various equivalent mathematical representations is
extremely important in solving the problem correctly.
We choose the Hamiltonian formulation as our basic form in practice based on the
fact that the Hamiltonian equations have symmetric and clean form, where the physical
laws of the motion can be easily represented. Secondly, the Hamiltonian formulation
is more general and universal than the Newton formulation. It can cover the classical,
relativistic, quantum, finite or infinite dimensional real physical processes where dis-
sipation effect can be neglected. Therefore, the success of the numerical methods for
Hamiltonian equations has broader development and application perspectives. Thus, it
is very surprising that the numerical algorithms for Hamiltonian equations are almost
nonexistent after we have searched various publications. This motivates us to study
the problem carefully to seek the answers to the previous two questions.
Our approach is to use the symplectic geometry, which is the geometry in phase
space. It is based on the anti-symmetric area metric, which is in contrast to the sym-
metric length metrics of Euclid and Riemann geometry. The basic theorem of the clas-
sic mechanics can be described as “the dynamic evolution of all Hamiltonian systems
preserves the symplectic metrics, which means it is a symplectic (canonical) transfor-
mation”. Hence the correct discretization algorithms to all the Hamiltonian systems
should be symplectic transformation. Such algorithms are called symplectic (canoni-
cal) algorithms or Hamiltonian algorithms. We have intentionally analyzed and eval-
uated the derivation of the Hamiltonian algorithm within the symplectic structures.
The fact proved that this approach is correct and fruitful. We have derived a series of
symplectic algorithms, found out their properties, laid out their theoretical foundation,
and tested them with extremely difficult numerical experiments.
In order to compare the symplectic and non-symplectic algorithm, we proposed
eight numerical experiments: harmonic oscillator, nonlinear Duffing oscillator, Huy-
gens oscillator, Cassini oscillator, two dimensional multi-crystal and semi-crystal lat-
tice steady flow, Lissajous image, geodesic flow on ellipsoidal surface, and Kepler
motion. The numerical experiments demonstrate the superiority of the symplectic al-
gorithm. All traditional non-symplectic algorithms fail without exception, especially
in preserving global property and structural property, and long-term tracking capabil-
ity, regardless of their accuracy. However, all the symplectic algorithms passed the
tests with long-term stable tracking capability. These tests clearly demonstrate the su-
periority of the symplectic algorithms.
Almost all of the traditional algorithms are non-symplectic with few exceptions.
They are designed for the asymptotic stable system which has dissipation mechanism
to maintain stability, whereas the Hamiltonian system does not have the asymptotic
stability. Hence all these algorithms inevitably contain artificial numerical dissipation,
fake attractors, and other parasitics effects of non-Hamiltonian system. All these ef-
fects lead to seriously twist and serious distortion in numerical results. They can be
used in short-term transient simulation, but are not suitable and can lead to wrong
Introduction 3
conclusions for long-term tracking and global structural property research. Since the
Newton equation is equivalent to Hamiltonian equation, the answer to the first ques-
tion is “No”, which is quite beyond expectation.
The symplectic algorithm does not have any artificial dissipation so that it can con-
genitally avoid all non-symplectic pollution and become a “clean” algorithm. Hamil-
tonian system has two types of conservation laws: one is the area invariance in phase
space, i.e., Liouville–Poincaré conservation law; the other is the motion invariance
which includes energy conservation, momentum and angular momentum conserva-
tion, etc. We have proved that all symplectic algorithms have their own invariance,
which has the same convergence to the original theoretical invariance as the conver-
gence order of the numerical algorithm. We have also proved that the majority of in-
variant tori of the near integrable system can be preserved, which is a new formulation
of the famous KAM (Kolmogorov–Arnorld–Moser) theorem[Kol54b,Kol54a,Arn63,Mos62] .
All of these results demonstrate that the structure of the discrete Hamiltonian algo-
rithm is completely parallel to the conservation law, and is very close to the original
form of the Hamiltonian system. Moreover, theoretically speaking, it has infinite long-
term tracking capability. Hence, a correct numerical method to solve the Newton equa-
tion is to Hamiltonize the equation first and then use the Hamiltonian algorithm. This
is the answer to the second question. We will describe in more detail the KAM theory
of symplectic algorithms for Hamiltonian systems in Chapter 13. In the following we
present some examples to compare the symplectic algorithm and other non-symplectic
algorithms in solving Newton equation of motion.
Calculation of the Harmonic oscillator’s elliptic orbit (Fig. 0.1(a)) uses Runge–Kutta
method (R–K) with a step size 0.4. The output is at 3,000 steps. It shows artificial
dissipation, shrinking of the orbit. Fig. 0.1(b) shows the results using Adams method
with a step size 0.2. It is anti-dissipative and the orbit is scattered out. Fig. 0.1(c)
shows the results of two-step central difference (leap-frog scheme). This scheme is
symplectic to linear equations. The results are obtained with a step size 0.1. It shows
that the results of three stages for 10,000,000 steps: the initial 1,000 steps, the middle
1,000 steps, and the final 1,000 steps. They are completely in agreement.
Fig. 0.2(a) shows the results of two-step central-difference. This scheme is non-
symplectic for nonlinear equations. The output is for step size 0.2 and 10,000 steps.
Fig. 0.2(a) shows the initial 1,000 steps and Fig. 0.2(b) shows the results between
9,000 to 10,000 steps. Both of them show the distortion of the orbit. Fig. 0.2(c) is for
the second-order symplectic algorithm with 0.1 step size, 1,000 steps.
4 Introduction
Using the R–K method, the two fixed points on the horizontal axes become two fake
attractors. The probability of the phase point close to the two attractors is the same.
The same initial point outside the separatrix is attracted randomly either to the left
or to the right. Fig. 0.3(a) shows the results with a step size 0.10000005 and 900,000
steps, which approach the left attractor. Fig. 0.3(b) shows the results with a step size
0.10000004 and 900,000 steps, which approach the right attractor. Fig. 0.3(c) shows
the results of the second-order symplectic algorithm with a step size 0.1. Four typical
orbits are plotted and each contains 100,000,000 steps: for every orbit first 500 steps,
the middle 500 steps, and the final 500 steps. They are in complete agreement.
(4) The dense orbit of the geodesic for the ellipsoidal surface
The dense orbit of the geodesic for the ellipsoidal surface with irrational frequency ra-
tio. The square of frequency ratio is 5/16, step size is 0.05658, 10,000 steps. Fig.0.4(a)
is for the R–K method which does not tend to dense. Fig. 0.4(b) is for the symplectic
algorithm which tends to dense.
6 Introduction
√
Fig. 0.4. Geodesics on ellipsoid, frequency ratio 5 : 4, non dense (a), dense orbit (b)
(5) The close orbit of the geodesic for the ellipsoidal surface
The close orbit of the geodesic for the ellipsoidal surface with rational frequency
ratio. The frequency ratio is 11/16, step size is 0.033427, 100,000 steps and 25 cycles.
Fig.0.5(a) is for the R–K method which does not lead to the close orbit. Fig. 0.5(b) is
for the symplectic algorithm which leads to the close orbit.
Fig. 0.5. Geodesics on ellipsoid, frequency ratio 11:16, non closed (a), closed orbit (b)
The close orbit of the Keplerian motion with rational frequency ratio. The frequency
ratio is 11/20, step size is 0.01605, 240,000 steps and 60 cycles. Fig. 0.6(a) is for the
R–K method which does not lead to the close orbit. Fig. 0.6(b)is for the symplectic
method which leads to the close orbit.
Introduction 7
Fig. 0.6. Geodesics on ellipsoid, frequency ratio 11:20, non closed (a), closed orbit (b)
d2 q ∂
m = − V,
d t2 ∂q
which is the standard formulation of the motion. It is a group of second-order differen-
tial equations in space Rn . It is usually called the standard formulation of the classical
mechanics, or Newton formulation.
Euler and Lagrange introduced an action on the difference between the kinetic
energy and potential energy
1
L(q, q̇) = T (q̇) − V (q) = (q̇, M q̇) − V (q).
2
Using the variational principle the above equation can be written as
d ∂L ∂L
− = 0,
d t ∂ q̇ ∂q
which is called the variational form of the classical mechanics, i.e., the Lagrange form.
In the 19th century, Hamilton proposed another formulation. He used the momen-
tum p = M q̇ and the total energy H = T + V to formulate the equation of motion
as
∂H ∂H
ṗ = − , q̇ = ,
∂q ∂p
which is called Hamiltonian canonical equations. This is a group of the first-order
differential equations in 2n phase space (p1 , · · · , pn , q1 , · · · , qn ). It has simple and
symmetric form.
8 Introduction
The three basic formulations of the classical mechanics have been described in
almost all text-books on theoretical physics or theoretical mechanics. These different
mathematical formulations describe the same physics law but provide different ap-
proaches in problem solving. Thus equivalent mathematical formulation can have
different effectiveness in computational methods. We have verified this in our own
simulations.
The first author did extensive research on Finite Element Method (FEM) in the
1960s [Fen65] which represents a systematic algorithm for solving equilibrium problem.
Physical problems of this type have two equivalent formulations: Newtonian, i.e., solv-
ing the second-order elliptic equations, and variational formulation, i.e., minimization
principle in energy functional. The key to the success of FEM in both theoretical and
computational methods lies in using a reasonable variational formulation as the basic
principle. After that, he had attempted to apply the FEM idea to the dynamic problem
of continuum media mechanics, but not yet achieved the corresponding success, which
appears to be difficult to accomplish even today. Therefore, the reasonable choice for
computational method of dynamic problem might be the Hamiltonian formulation.
Initially it is a conjecture and requires verification from the computational experi-
ments. We have investigated how others evaluated the Hamiltonian system in history.
First we should point out that Hamilton himself proposed his theory based on the ge-
ometric optics and then extended it to mechanics that appears to be a very different
field. In 1834 Hamilton said, “This set of idea and method has been applied to optics
and mechanics. It seems it can be applied to other areas and developed into an inde-
pendent knowledge by the mathematicians”[Ham34] . This is just his expectation, and
other peers in the same generation seemed indifferent to this set of theory, which was
“beautiful but useless”[Syn44] to them. Klein, a famous mathematician, while giving a
high appreciation to the mathematical elegance of the theory, suspected its applicabil-
ity, and said: “. . . a physicist, for his problems, can extract from these theories only
very little, and an engineer nothing”[Kle26] . This claim has been proved wrong at least
in physics aspect in the later history. The quantum mechanics developed in the 1920s
under the framework of the Hamiltonian formulation. One of the founders of the quan-
tum mechanics, Schrödinger said, “Hamiltonian principle has been the foundation for
modern physics . . . If you want to solve any physics problem using the modern theory,
you must represent it using the Hamiltonian formulation”[Sch44] .
The Hamiltonian system is one of the most important systems among all the dynam-
ics systems. All real physical processes where the dissipation can be neglected can be
formulated as Hamiltonian system. Hamiltonian system has broad applications, which
include but are not limited to the structural biology, pharmacology, semiconductivity,
superconductivity, plasma physics, celestial mechanics, material mechanics, and par-
tial differential equations. The first five topics have been listed as “Grand Challenges”
in Research Project of American government.
Introduction 9
The development of the physics verifies the importance of the Hamiltonian sys-
tems. Up to date, it is undisputed that all real physical processes where the dissipation
can be neglected can be written as Hamiltonian formulation, whether they have finite
or infinite degrees of freedom.
The problem with finite degrees of freedom includes celestial and man-made
satellite mechanics, rigid body, and multi-body (including the robots), geometric op-
tics, and geometric asymptotic method (including ray-tracing approximation method
in wave-equation, and WKB equation of quantum mechanics), confinement of the
plasma, the design of the high speed accelerator, automatic control, etc.
The problem with infinite degrees of freedom includes ideal fluid dynamics, elas-
tic mechanics, electrical mechanics, quantum mechanics and field theory, general rel-
ativistic theory, solitons and nonlinear waves, etc.
All the above examples show the ubiquitous and nature of the Hamiltonian sys-
tems. It has the advantage that different physics laws can be represented by the same
mathematical formulation. Thus we have confidence to say that successful develop-
ment of the numerical methods for Hamiltonian system will have extremely broad
applications.
We now discuss the status of the numerical method for Hamiltonian systems.
Hamiltonian systems, including finite and infinite dimensions, are Ordinary Differ-
ential Equations (ODE) or Partial Differential Equations (PDE) with special form.
The research on the numerical method of the differential equations started in the 18th
century and produced abundant publications. However, we find that few of them dis-
cuss the numerical method specifically for Hamiltonian systems. This status is in sharp
contrast with the importance of the Hamiltonian system. Therefore, it is appealing and
worthy to investigate and develop numerical methods for this virgin field.
which is the area of the parallel quadrilateral with vectors x and y as edges. Generally
speaking, the symplectic inner product is an area metric. Due to the anti-symmetry
of the inner product, [x, x] = 0 always holds for any vector x. Thus it is impos-
sible to derive the concept of length of a vector from the symplectic inner product.
This is the fundamental difference between the symplectic geometry and Euclid ge-
ometry. All transformations that preserve the symplectic inner product form a group,
called a symplectic group, Sp(2n), which is also a typical Lie group. Its corresponding
Lie algebra consists of all infinitesimal symplectic transformations B, which satisfy
B T J + JB = 0. We denote it as sp(2n). Since the non-degenerate anti-symmetric
matrix exists only for even dimensions, the symplectic space must be of even dimen-
sions. The phase space exactly satisfies this condition.
Overall the Euclid geometry is a geometry for studying the length, while the sym-
plectic geometry is for studying the area.
The one-to-one nonlinear transformation in the symplectic geometry is called sym-
plectic transformation, or canonical transformation. The transformation whose Jaco-
bian is always a symplectic matrix plays a major role in the symplectic geometry. For
the Hamiltonian system, if we represent a pair of n-dim vectors with a 2n-dim vector
z = (p, q), the Hamiltonian equation becomes
dz ∂H
= J −1 .
dt ∂z
Under the symplectic transformation, the canonical form of the Hamiltonian equation
is invariant. The basic principle of the Hamiltonian mechanics is for any Hamiltonian
system. There exists a group of symplectic transformation (i.e., the phase flow) GtH1 ,t0
that depends on H and time t0 , t1 , so that
which means that GtH1 ,t0 transforms the state at t = t0 to the state at t = t1 . Therefore,
all evolutions of the Hamiltonian system are also evolutions of the symplectic trans-
formation. This is a general mathematical principle for classical mechanics. When
H is independent of t, GtH1 ,−t2 = GtH1 ,−t0 , i.e., the phase flow depends only on the
difference in parameters t1 − t0 . We can let GtH = Gt,0 H .
One of the most important issues for the Hamiltonian system is stability. The fea-
ture of this type of problems in geometry perspective is that its solution preserves the
metrics. Thus the eigenvalue is always a purely imaginary number. Therefore, we can-
not use the asymptotic stability theory of Poincaré and Liapunov. The KAM theorem
must be used. This is a theory about the total stability and is the most important break-
through for Newton mechanics. The application of the symplectic geometry to the nu-
merical analysis was first proposed by K. Feng [Fen85] in 1984 at the international con-
ference on differential geometry and equations held in Beijing. It is based on a basic
principle of the analytical mechanics: the solution of the system is a volume-preserved
transformation (i.e., symplectic transformation) with one-parameter2 on symplectic
2
Before K.Feng’s work, there existed works of de Vorgelaere[Vog56] , Ruth[Rut83] and
Menyuk[Men84] .
12 Introduction
integration. Since then, new computational methods for the Hamiltonian system have
been developed and we have studied the numerical method of the Hamiltonian system
from this perspective. The new methods make the discretized equations preserve the
symplectic structure of the original system, i.e., to restore the original principle of the
discretized Hamiltonian mechanics. Its discretized phase flow can be regarded as a
series of discrete symplectic transformations, which preserve a series of phase area
and phase volume. In 1988, K. Feng described his research work on the symplectic
algorithm during his visit to Western Europe and gained the recognition from many
prominent mathematicians. His presentation on “Symplectic Geometry and Compu-
tational Hamiltonian Mechanics” has obtained consistent high praise at the workshop
to celebrate the 60th birthday of famous French mathematician Lions. Lions thought
that K. Feng founded the symplectic algorithm for Hamiltonian system after he devel-
oped the finite element methods independent of the efforts in the West. The prominent
German numerical mathematician Stoer said, “This is a new method that has been
overlooked for a long time but should not be overlooked.”
We know that we can not study the Hamiltonian mechanics without the symplectic
geometry. In the meantime, the computational method of the Hamiltonian mechanics
doesn’t work without the symplectic difference scheme. The classical R–K method is
not suitable to solve this type of problems, because it cannot preserve the long-term
stability. For example, the fourth-order R–K method obtains a completely distorted
result after 200,000 steps with a step size 0.1, because it is not a symplectic algorithm,
but a dissipative algorithm.
We will describe in more detail the theory of symplectic geometry and symplectic
algebra in Chapters 1, 2 and 3.
In addition, the first author and his group constructed various symplectic schemes
with arbitrary order of accuracy using the generating function theory from the ana-
lytical mechanics perspective. In the meantime, he extended the generating function
theory and Hamilton–Jacobi equations by constructing all types of generating function
and the corresponding Hamilton–Jacobi equations. The generating function theory and
the construction of the symplectic schemes will be introduced in Chapter 5.
Among the various dynamical systems, one of them is called source-free dynamical
system, where the divergence of the vector field is zero:
dx
= f (x), div f (x) = 0.
dt
∂xn+1
The phase flow to this system is volume-preserved, i.e., det = 1. Therefore,
∂xn
the numerical solution should also be volume-preserved.
We know that Hamiltonian system is of even dimensions. However, the source-free
system can be of either even or odd dimensions. For the system of odd dimensions,
the Euler midpoint scheme may not be volume-preserved. ABC (Arnold–Beltrami–
Childress) flow is one of the examples. Its vector field has the following form:
ẋ = A sin x + C cos y,
ẏ = B sin x + A cos z,
ż = C sin y + B cos x,
which is a source-free system and the phase flow is volume-preserved. This is a split
system and constructing the volume-preserving scheme is easy. Numerical experi-
ments show that the volume-preserving scheme can calculate the topological structure
accurately, whereas the traditional schemes can not[FS95,QZ93] . We will give more de-
tails in Chapter 10.
There exists a special type of dynamical systems with odd dimensions. They have
similar symplectic structure as the systems of even dimensions. We call them contact
systems. The reader can find more details in Chapter 11.
Consider the contact system in R2n+1 space
14 Introduction
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
x x1 y1
⎢ ⎥ ⎢ ⎥
(2n + 1) − dim vector : ⎣ y ⎦ , where x = ⎣ ... ⎦ , y = ⎣ ... ⎦ , z = (z);
z xn y
⎡ ⎤ ⎡ ⎤ ⎡n ⎤
a(x, y, z) a1 b1
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
(2n + 1) − dim v.f. : ⎣ b(x, y, z) ⎦ , where a = ⎣ ... ⎦ , b = ⎣ ... ⎦ , c = (c).
c(x, y, z) an bn
A contact system can be generated from a contact Hamiltonian function K(x, y, z):
dx
= −Ky + Kz x = a,
dt
dy
= Kx = b,
dt
dz
= Ke = c,
dt
Ke (x, y, z) = K(x, y, z) − (x, Ky (x, y, z)).
⎡ ⎤
dx
α = xd y + d z = [0, x, 1] ⎣ d y ⎦ .
dz
reasonable. Furthermore, because the errors of the energy are controlled, the errors of
numerical trajectories of celestial bodies are no longer along the track by (t − t0 )2
laws of the fast-growing, and with only a t − t0 linear growth, this to the long arc
computation is extremely advantageous.
For the advantages of the symplectic algorithms, nowadays they have been widely
used in the study of dynamical astronomy, especially in the qualitative analysis of the
evolution of solar system, e.g. to analyze the stable motion area, space distributions
and trajectory resonance of little planets, long time evolution of large planets and
extra-planets, and other hot topics in the dynamical astronomy.
We first use two simple examples to illustrate the special affects of symplectic schemes
on the qualitative analysis in dynamics astronomy[LZL93,JLL02,LL95,LL94,Lia97,LLZW94] .
Example I. The Keplerian motions. It is the elliptical motions of two-body problem.
The corresponding Hamiltonian function is:
where p and q are the generalized coordinates and generalize momentum, T and V
are the kinetic and potential energies. The analytic solution is a fixed ellipse. When
we simulate this problem by the R–K methods and symplectic algorithms, the former
ones shrink the ellipse gradually, whereas the later ones preserve the shape and size
of the ellipse (see the numerical trajectories after 150 and 1000 steps respectively in
Fig. 0.7(a), e=0.7 and Fig. 0.7(b), e=0.9 where e is the eccentricity of the ellipse).
This means the non-symplectic R–K methods have the false energy dissipation and
the symplectic algorithms preserve the main character of the Kepler problem because
of the conservation of the symplectic structure.
Example II. The axial symmetry galaxy’s stellar motion question. Its simplified
dynamic model corresponding to the Hamiltonian function is:
1 2 1 2
H(p, q) = (p1 + p22 ) + (q12 + q22 ) + (2q12 q2 − p32 ).
2 2 3
To obtain the basic character of the dynamics of this system, we compute it with
order 7 and order 8 Runge–Kutta–Fehlberg methods (denoted as RKF(7) and RKF(8)
resp.), as well as the order 6 explicit symplectic algorithm (SY6). The numerical re-
sults are listed in Fig. 0.8 to Fig. 0.10. In these figures (Fig. 0.8 to Fig. 0.9), we see that
the symplectic algorithm preserves the energy H very well in both of the two cases
(ordered LCN= 0 and disorder region LCN> 0), while the RKF methods increase the
energy with the evolution of time ΔH. In Fig. 0.10 (a) and Fig. 0.10 (b), the symplec-
tic algorithms present numerically the basic characters of the system: the fixed curve
in case of LCN= 0 and the chaos property in case of LCN> 0.
Introduction 17
Fig. 0.7. Comparison of calculation of Keplerian motion by R–K and symplectic methods.
Fig. 0.8. Curves of ΔH obtained by RKF(8)[left] and SY6 [right] both with H0 = 0.553,
LCN=0.
Fig. 0.9. Curves of ΔH obtained by RKF(8)[left] and SY6 [right] both with H0 =
0.0148, LCN > 0
Fig. 0.10. Poincaré section obtained by RKF(8)[left] with H0 = 0.553,LCN=0 and SY6
[right] with H0 = 0.0148,LCN>0
the symplectic algorithms were widely used to study the dynamical astronomy. Cur-
rently, it is a hot topic to study the dynamical evolution of the solar system, such
as the long-term trajectory evolution of large planet and extra-planet, the space dis-
tribution of little planets in main zone (Kirkwood interstice phenomenon), trajectory
resonance, the evolution of satellite system of a large planet, the birth and evolution of
planet loops and the trajectory evolution of a little planet near Earth. All these prob-
lems require numerical simulation for a very long time, e.g. 109 years or more for the
solar system. Thus, the time steps for the numerical methods shall be large enough
due to limitations of our computers, while the basic property of the system should be
preserved. This excludes all the non-symplectic methods, whilst just lower order sym-
plectic algorithms are valid for the task. In recent years, many astronomers in Japan
and America, e.g. Kinoshita[KYN91] , Bretit[GDC91] and Wisdom[WHT96,WH91] , have done
a large amount of research on the evolution of the solar system. The following con-
tribution of Wisdom has been widely cited. He derived the Hamiltonian function in
Jacobin coordinates of the solar system as
n−1
H(p, q) = Hi (p, q) + εΔH(p, q),
i=1
1◦ For the restrictive three-body system constituted by solar, the major planet
and the planetoid, some new results have been obtained after studying its correspond-
ing resonance of 1:1 orbit and the triangle libration point. These results can success-
fully explain the distribution of stability region [ZL94,ZLL92] of Trojan planetoid, as well
the actual size of the stable region of distributed triangle libration points corresponding
to several relate major planet.
2◦ Adopting the splitting method of Wisdom for the Hamiltonian function to
study the long-term trajectories evolution of some little planets.
two problems, respectively. The numerical results of the errors Δ(M + ω) of main
1000 cycles trajectory are listed in Table 0.1 and Table 0.2. From the two tables, we
can clearly see that the errors of the non-symplectic methods, though very small at the
beginning, increase rapidly as (t − t0 )2 ; whereas the errors of symplectic algorithm
increase linearly as t − t0 . The results of symplectic algorithms are much better. This
indicates that though the accuracy order of symplectic algorithms is the same as for
other methods, they have more application value in the quantitative computations. We
also improve the RKF7(8) for energy conserving methods by compensating the en-
ergy at every time step. We denote such method as the RKH method whose numerical
results are also listed in the two tables. From the results, we can see that we have
made much improvement of the schemes. The results of the energy error by the RKH
are almost same with those by the symplectic algorithm. Thus the RKH methods not
only have high order accuracy, but also can preserve the energy approximately as the
symplectic algorithms.
Every operator is unitary and depends on t1 , t2 and Ĥ. They are independent of the
state ψ(t1 , r) at time t1 . Therefore, the time evolutions of the quantum system are
evolutions of unit transformation in this sense. Every operator can induce an operator,
which acts on the real function vector. The two components of the real functions vector
are the real part and the image part of the wave function, i.e.,
b(t , r) b(t , r)
2 t1 ,t2 1
= SĤ .
a(t2 , r) a(t1 , r)
t1 ,t2
The operator SĤ preserves the inner product and symplectic wedge product for
any two real function vectors. It is simply called norm-preserving symplectic evolu-
tion. The quantum system is a Hamiltonian system (with infinite dimensions) and the
time evolution of the Schrödinger equation can be rewritten as a canonical Hamilto-
nian system for the two real functions of the wave function as the generalized mo-
mentum and generalized coordinates. The norm of wave function is the conservation
law of the canonical system. Thus it is reasonable to integrate such a system by the
norm-preserving symplectic numerical methods. To apply such a method to the in-
finite dimensional system, we should first space discretize the system into a finite
dimensional canonical Hamiltonian system, which also preserves the norm of wave
function. Suppose the characteristic functions of the operator Ĥ0 (r) for the evolution-
ary Schrödinger equation with some given boundary conditions contain the discrete
states and continuous states.
When the Hamiltonian Ĥ is independent on time explicitly, the energy of the quan-
tum system ψ|, Ĥ|ψ = Z T HZ is a conservation law both for the canonical system
and norm-preserving symplectic algorithm. Such a norm-preserving symplectic algo-
rithm with the fourth order accuracy can be constructed by the order 4 diagonal Padé
approximations to the exponential function eλ .
In the following, we take an example to introduce the method to discretize the time
involved Schrödinger equation to a canonical system[LQHD07,QZ90a] .
Consider the time evolution of an atom moving in one dimensional space by the
action of some strong field V (t, x) is
∂ψ
i = Ĥψ, Ĥ = Ĥ0 (r) + V̂ (t, r),
∂t
1 ∂2
Ĥ0 = − + V0 (x),
2 ∂ x2
0, 0 < x < 1,
V0 (x) =
∞, x ≤ 0 or x ≥ 1.
22 Introduction
In contrast to the characteristic function expanding method, we don’t make any trun-
cation for the wave function when discretizing the Schrödinger equation. Therefore,
the resulting canonical system contains all the characteristic states of Ĥ0 .
The numerical conservation laws of explicit symplectic algorithms will converge
to the corresponding conservation laws of the system as the time step tends to zero.
Thus, although numerical energy and norm of the wave function presented by explicit
symplectic algorithms will not be preserved exactly, they will converge to the true
energy and norm of the wave function of the system as the time step reduces.
The time dependent Schrödinger equation (TDSE) in one dimensional space by
the action of some strong field V (t, x) is
∂ψ
i = Ĥψ, Ĥ = Ĥ0 (x) + εV̂ (t, x),
∂t
1 ∂2
Ĥ0 = − + V0 (x),
2 ∂x2
0, 0 < x < 1,
V0 (x) =
∞, x ≤ 0 or x ≥ 1.
⎧
⎪
⎨ 2x, 0 < x < 0.5,
V (x) = 2x − 2x, 0.5 ≤ x ≤ 1,
⎪
⎩ 0, x ≤ 0 or x ≥ 1.
n2 π 2
S = (Smn ), Smn = δmn + εvmn ,
⎧ 2
⎪ 1 1 − (−1)n
⎪
⎪ + , m = n,
⎪
⎪ 2 n2 π 2
⎪
⎪ |m − n| = 1, 3, 5, · · · ,
⎪
⎨ 0, m−n
vmn = −16mn(1 − (−1) 2 )
⎪
⎪ , |m − n| = 2, 4, 6, · · · , n = 2, 4, 6, · · · ,
⎪
⎪ (m2 − n2 )2 π 2
⎪
⎪ m−n
⎪
⎪ −8|2mn − (−1) 2 (m2 − n2 )|
⎩ , |m − n| = 2, 4, 6, · · · , n = 1, 3, 5, · · · .
2 (m − n ) π
2 2 2
1◦ The R–K method can not preserve the energy and the norm of wave function,
as evident by ER–K in Fig. 0.11(left) and NR–K in Fig. 0.11(right).
2◦ The Euler midpoint rule can preserve the energy and norm, as evident by EE
in Fig. 0.11(left) and NE in Fig. 0.11(right). Note that for EE in Fig. 0.11(left), there
is a very small increase at some time because of the implicity of the Euler scheme.
Fig. 0.11. Energy [left] and norm [right] comparison among the 3 difference schemes
3◦ The explicit symplectic algorithms can preserve exactly the energy Ẽ(bk , ak ; h)
and norm Ñ (bk , ak ; h), as evident by ES in Fig. 0.11(left) and NS in Fig. 0.11(right).
If we want to get further insight into these conservation laws within smaller scales, we
find that as the time steps get smaller, the numerical energy of symplectic algorithm
converges to the true energy of the system e0 = 42.0110165 and the numerical norm
converges to unit n0 = 1. See Table 0.3 showing the numerical energy and norm as
well as their errors. The errors are defined as
Actually, the numerical energy and norm obtained by symplectic algorithm oscillate
slightly, as shown by ES and NS in Fig. 0.12. However, the amplitude of their oscil-
lations will converge to zero, if the time step tends to zero. As the time step tends to
zero, we have
e(h) −→ e0 , CE (h) = maxk |ESk − e0 | −→ 0,
Table 0.3. The change of energy and norm of the wave function with the step size
Fig. 0.12. Energy E and norm N obtained from explicit symplectic scheme
In all, for a quantum system with real Hamiltonian function independent of time
explicitly, the explicit symplectic algorithms can preserve the energy and norm of the
wave function to any given accuracy. They overcome the main disadvantages of the
traditional numerical methods.
Next, we look at the quantum system with real Hamiltonian function, which is
dependent on time explicitly. In this case, the resulting system after semi-discretization
is an m-dimensional, separable, linear, Hamiltonian canonical system. The energy of
the system is not conserved any more, but the norm of the wave function is still a
quadratic conservation law.
The TDSE for an atom in one dimensional space with the action of some strong
field V (t, x) = εx sin(ωt) is
∂ψ
i = Ĥψ, Ĥ = Ĥ0 (x) + εV̂ (t, x),
∂t
1 ∂2
Ĥ0 = − + V0 (x).
2 ∂ x2
By the similar method
√ as before, we expand the wave function as the characteristic
functions Xn x = 2 sin nπx(n = 1, 2, · · ·) of Ĥ0 to discretize the TDSE. Because
Introduction 25
the Hamiltonian operator is real, the discrete TDSE is a separable linear canonical
Hamiltonian system with the parameters as follows.
n2 π 2
S(t) = (s(t)mn ), s(t)mn = δm,n + εv(t)mn ;
2
⎧
⎪
⎪ sin(ωt), m = n,
⎪
⎨
v(t)mn = 0, |m − n| = 2, 4, 6, · · · ,
⎪
⎪
⎪
⎩ 8mn sin (ω t) , |m − n| = 1, 3, 5, · · · .
(m2 − n2 )2 π 2
√
The initial state is taken as ψ(0, x) = X1 (x) = 2 sin(πx). The energy of the
system is not conserved in this case because the Hamiltonian depends on the time
explicitly. The norm of wave function remains unitary, i.e., N (b, a) = n0 = 1. We
take the Euler midpoint rule scheme, order 2 explicit symplectic algorithm and the
order 2 R–K method to compute the problem with the same time step h = 4 × 10−3 .
The numerical results are as follows:
1◦ The R–K method increases the norm of wave function rapidly, see NR–K in
Fig. 0.13(left). It leads to unreasonable results, see in Fig. 0.13(right).
2◦ The Euler midpoint rule scheme can preserve the norm, see NE in Fig.0.13(left).
These results are in good agreement with the theoretical results. See Fig. 0.13(right)
π
for the results for weak fields ε = . When ω = ΔE1n , i.e., resonance occurs, the
2
basic state and the first inspired state will intermix and the variation period of the
energy is identical to the period of intermixing. See the corresponding results in Fig.
0.14(left) and Fig. 0.14(right). When ω = ΔE1n there will not be intermixing. See the
corresponding numerical results in Fig. 0.15(left) and Fig. 0.15(right), where O is the
basic state. When the field is strong, the selection rule is untenable, and no resonance
occurs, but the basic state will intermix with the first, second, . . . inspired states. See
5π 2 3π 2
the results for ω = in Fig. 0.16(left) and Fig. 0.16(right) and ω = = ΔE12
4 2
in Fig. 0.17(left) and Fig. 0.17(right).
26 Introduction
3◦ The order 2 explicit symplectic algorithms can not preserve the norm exactly.
The numerical norms oscillate near the unit. See NS in Fig. 0.13, where changes of
numerical energy and states of intermixing obtained by symplectic algorithms are
similar to the results of Euler midpoint rule scheme.
We can conclude that for this system the R–K method can not preserve the norm of
wave function and its results are unreasonable; the Euler scheme can preserve the norm
and its results are in agreement with the theoretical results; the second order scheme
obtains the numerical norm which oscillates near the unit and its energy and states of
intermixing are the same as for the results of Euler scheme. Thus, the Euler scheme
(an implicit symplectic scheme) and the second order explicit symplectic algorithm
are good choices for studying the quantum system with the Hamiltonian dependent on
time explicitly. They overcome the drawbacks of the traditional R–K methods.
The classical trajectory method regards the atom approximatively as a point and
the system as a system of some points, and advances the process of action as the clas-
sical motions of point system in potential energy plane of the electrons. It was Bunker
who first applied the R–K method to computations of classical trajectory of molecular
reacting system. Karplus et al. did a large number of computations by all kinds of
numerical methods and screened out the R–K–G (Runge–Kutta–Gear) method to pro-
long the computation time from 10−15 s to 10−12 s. The R–K–G method made rapid
progress in the theoretical study of reacting dynamics of microscopic chemistry and
was widely used for computation of classical trajectory. However, its valid computa-
tion time is much less than 10−8 s which is necessary time for study of chemical re-
actions. Moreover, there were many differences between the numerical quantities and
theoretical quantities of some parameters. The classical trajectory method describes
the microscopic reaction system approximately as a Hamiltonian system which natu-
rally has symplectic structure. Thus, it is expected that the symplectic algorithms will
overcome the shortages of the R–K–G method and improve the numerical results.
Here we take the mass of the proton as the unit mass and 4.45 × 10−14 s as unit
time.
Consider the classical motions of the A2 B type molecules like H2 O and SO2 mov-
ing in the electron potential energy plane of the reaction system and preserving the
28 Introduction
q1 = z1 + z2 − 2z3 , q2 = y 2 − y 1 ,
and the generalized mass as M1 = 0.25, M2 = 0.5, further the generalized momen-
tum as
d q1 d q2
p1 = 0.25 , p2 = 0.5 ,
dt dt
and the kinetic energy of system as
The potential energy suggested by Banerjee, who introduced the symmetry C2v
and notation D = q12 + q22 , was
and further its classical trajectories of A2 B system and the changes of kinetic energy,
potential energy and total energy with time by following relations:
q1 q2 q1
y3 = 0, z3 = − ; y2 = −y1 = , z2 = z1 = .
4 2 4
Fig. 0.18. The potential energy curve of the electronic potential function in phase space
We compute this system with order 4 explicit symplectic algorithm and R–K
method. The time step is taken as h = 0.01 for both. The numerical classical tra-
jectories, kinetic energy, potential energy and total energy are recorded. Fig. 0.18
shows the potential energy curve of the electronic potential function in phase space. If
|q1 | → +∞, then V (q) → +∞; if |q2 | → 0 or |q2 | → +∞, then V (q) → +∞. By
the theoretical analysis, we know that the total energy of the system will be conserved
all the time, the three atoms will oscillate nearly periodically, and the whole geometry
structure of the system may be reversed but kept periodic. The changes of the total
energy with time are shown in Fig.0.19, where we can see that the total energy ob-
tained by symplectic algorithms are preserved up to 6.23 × 10−9 s, whereas the R–K
method reduces them rapidly with time. The motion trajectories of the system in the
plane by the symplectic algorithms and R–K method are shown in Fig. 0.20 (a), (c),
(e) and (b), (d), (f) resp., where we can see that the numerical results of symplectic al-
gorithms are coincident with the theoretical results but the results of R–K method are
not. We also applied the order 1 and 2 symplectic algorithms, the Euler method and the
revised Euler method to compute the same problem. The conclusions are almost the
same. Because all the traditional methods such as R–K methods, Adams methods and
Euler methods can not preserve the symplectic structure of this microscopic system,
they will bring false dissipations inevitably, which will make their numerical results
meaningless after long-term computations. On the contrary, symplectic algorithms can
preserve the structure and do not bring any false dissipations. Therefore, they are suit-
able for long-term computations and greatly improve the classical trajectory methods
for studying the microscopic dynamical reactions of chemical systems.
30 Introduction
Fig. 0.19. The changes of the total energy with the time
dp d V (q)
=− = −f (q),
dt dt
dq d U (p)
= = g(p).
dt dt
Fig. 0.20. The motion trajectories of the system in the plane,(a) and (b) period range from
4.45 × 10−10 s to (4.45 × 10−10 + 4.45 × 10−13 )s. (c) and (d) period range from 6.23 × 10−9 s
to (6.23 × 10−9 + 4.45 × 10−13 )s. (e) and (f) period range from 6.23 × 10−9 s to (6.23 ×
10−9 + 4.45 × 10−13 )s. (a), (c), (e) is the symplectic algorithm path, (b), (d), (f) is the R–K
method path
as well as the changes of kinetic energy, potential energy and total energy with the
variation of time.
We compute some states of two homonuclear molecules Li2 and N2 and two
heteronuclear molecules CO and CN by using the order 1, 2 and 4 explicit sym-
plectic algorithms and compare the numerical results of total energy and classical
trajectories with the Euler method and order 2 and 4 order R–K methods. In Fig.
0.21, Fig. 0.22 and Fig. 0.23, we show the numerical results of the classical tra-
jectories, total energy and the trajectories in p − q phase space obtained by order
4 explicit symplectic algorithm and order 4 R–K method respectively. The parame-
32 Introduction
Recently, the symplectic algorithms have been applied to study the observation opera-
tor of the global positioning system (GPS) by Institute of Atmospheric Physics of the
Chinese Academy of Science[WZJ95,WJX01] . Numerical weather forecasting needs very
large amount of atmospheric information from GPS. One of the key problems in this
field is how to reduce largely the computational costs and to compute it accurately
for a long time. The symplectic algorithms provide rapid and accurate numerical al-
gorithms for them to deal with the information of GPS efficiently. The computational
costs of the symplectic algorithms are one four hundredth of the costs of traditional
algorithms. For the complicated nonlinear system of atmosphere and ocean, symplec-
tic algorithms can preserve its total energy, total mass, total potential so well that the
relative errors of potential height is below 0.0006 (see Fig. 0.24).
Another application of symplectic algorithms to geophysics is carried out by In-
stitute of Geophysics to prospect for the oil and natural gas[GLCY00,LLL01a,LLL01b,LLL99] ,
which has obtained several great achievements. For example, the spread waves of
earthquake under the framework of Hamiltonian system and the corresponding sym-
plectic algorithms have been investigated. Moreover, “the information of oil reserves
and geophysics and its process system ” has been produced, and the task of prospecting
for 1010 m3 of natural gas, which has obtained. Fig. 0.25 shows the numerical results of
prestack depth migration in the area of Daqing Xujiaweizi by applying symplectic al-
gorithms to Marmousi model. Recently, Liuhong et.al. proposed a new method[LYC06]
to calculate the depth extrapolation operator via exponential of pseudo-differential op-
erator in lateral varied medium. The method offers the phase of depth extrapolation
operator by introducing lateral differential to velocity, which in fact is an application
of Lie group method.
34 Introduction
Fig. 0.24. The relative errors of potential height is below 0.0006 after 66.5 days
Fig. 0.25. Numerical results of prestack depth migration in the area of Daqing Xujiaweizi
obtained by applying symplectic algorithms to Marmousi model
Bibliography
[FW94] K. Feng and D.L. Wang: Dynamical systems and geometric construction of algo-
rithms. In Z. C. Shi and C. C. Yang, editors, Computational Mathematics in China, Con-
temporary Mathematics of AMS Vol 163, pages 1–32. AMS, (1994).
[GDC91] B. Gladman, M. Duncan, and J. Candy: Symplectic integrators for long-term inte-
gration in celestial mechanics. Celest. Mech., 52:221–240, (1991).
[GLCY00] L. Gao, Y. Li, X. Chen, and H. Yang: An attempt to seismic ray tracing with
symplectic algorithm. Chinese Journal Geophys., 43(3):402–409, (2000).
[Gol80] H. Goldstein: Classical Mechanics. Addison-Wesley Reading, Massachusetts, (1980).
[GS84] V. Guillemin and S. Sternberg: Symplectic Techniques in Physics. Cambridge Univer-
sity Press, Cambridge, (1984).
[GS94a] Z. Ge and C. Scovel: Hamiltonian truncation of shallow water equations. Letters in
Mathematical Physics, 31:1–13, (1994).
[Ham34] Sir W. R. Hamilton: On a general method in dynamics; by which the study of the
motions of all free systems of attracting or repelling points is reduced to the search and
differentiation of one central relation, or characteristic function. Phil. Trans. Roy. Soc.
Part II for 1834, 247–308; Math. Papers, Vol. II, 103–161, Second edition, (1834).
[Ham40] W.R. Hamilton: General methods in dynamics, volume I,II. Cambridge Univ. Press,
(1940).
[HIWZ95] T. Y. Huang, K. A. Innanen, C. B. Wang, and Z. Y. Zhao: Symplectic methods
and their application to the motion of small bodies in the solar system. Earth, Moon, and
Planets,, 71(3):179–183, (1995).
[HKRS97] M. Hankel, B. Karasözen, P. Rentrop, and U. Schmitt: A Molecular Dynamics
Model for Symplectic Integrators. Mathematical Modelling of Systems, 3(4):282–296,
(1997).
[HL97a] E. Hairer and P. Leone: Order barriers for symplectic multi-value methods. In D.F.
Grifysis, D.F.Higham, and G.A. Watson, editors, Numerical analysis 1997 Proc. of the 17th
Dundee Biennial Conference, June 24-27, 1997, Pitman Reserch Notes in math. series 380,
pages 133–149, (1997).
[IMKNZ00] A. Iserles, H. Z. Munthe-Kaas, S. P. Nørsett, and A. Zanna: Lie-group methods.
Acta Numerica, 9:215–365, (2000).
[JLL02] J. Ji, G. Li, and L. Liu: The dynamical simulations of the planets orbiting gj 876.
Astrophys. J., 572: 1041C-1047, (2002).
[Kle26] F. Klein: Vorlesungen über die Entwicklung der Mathematik in 19 Jahrhundert. Teub-
ner, (1926).
[Kol54a] A. N. Kolmogorov: General theory of dynamical systems and classical mechanics.
In Proc. Inter. Congr. Math., volume 1, pages 315–333, (1954).
[Kol54b] A. N. Kolmogorov: On conservation of conditionally periodic motions under small
perturbations of the Hamiltonian. Dokl. Akad. Nauk SSSR, 98:527–530, (1954).
[KYN91] H. Kinoshita, H. Yoshida, and H. Nakai: Symplectic integrators and their application
to dynamical astronomy. Celest. Mech. and Dyn. Astro., 50:59–71, (1991).
[LDJW00] Y. X. Li, P. Z. Ding, M. X. Jin, and C. X. Wu: Computing classical trajectories of
model molecule A2 B by symplectic algorithm. Chemical Journal of Chinese Universities,
15(8):1181–1186, (2000).
[Lia97] X. Liao: Symplectic integrator for general near-integrable Hamiltonian systems. Ce-
lest. Mech. and Dyn. Astro., 66:243–253, (1997).
[LL94] L. Liu and X. H. Liao: Numerical calculations in the orbital determination of an artifi-
cial satellite for a long arc. Celest. Mech., 59:221–235, (1994).
[LL95] L. Liu and X. H. Liao: Existence of formal integrals of symplectic integrators. Celest.
Mech., 63(1):113–123, (1995).
[LL99] L. D. Landau and E. M. Lifshitz: Mechanics, Volume I of Course of Theoretical
Physics. Corp. Butterworth, Heinemann, New York, Third edition, (1999).
[LLL99] M. Luo, Y. Li, and H. Lin: The symplectic geometric description and algorithm of
seismic wave propagation. In The 69-th Ann. Seg mecting, volume 199, pages 1825–1828,
(1999).
Bibliography 37
[LLL01a] Y. M. Li, H. Liu, and M. Q. Luo: Seismic wave modeling with implicit symplectic
method based on spectral factorization on helix. Chinese Journal Geophys., 44(3):379–388,
(2001).
[LLL01b] M. Q. Luo, H. Liu, and Y. M. Li: Hamiltonian description and symplectic method
of seismic wave propagation. Chinese Journal Geophys., 44(1):120–128, (2001).
[LLZD01] X. S. Liu, X. M. Liu, Z. Y. Zhao, and P. Z. Ding: Numerical solution of 2-D time-
independent schrödings. Int. J. Quant. Chem., 83:303–309, (2001).
[LLZW94] L. Liu, X. Liao, Z. Zhao, and C. Wang: Application of symplectic integrators to
dynamical astronomy(3). Acta Astronomica Sinica, 35:1, (1994).
[LQ88] C.W. Li and M.Z. Qin: A symplectic difference scheme for the infinite dimensional
Hamiltonian system. J. Comput. Appl. Math., 6:164–174, (1988).
[LQ95a] S. T. Li and M. Qin: Lie–Poisson integration for rigid body dynamics. Computers
Math. Applic., 30:105–118, (1995).
[LQ95b] S. T. Li and M. Qin: A note for Lie–Poisson Hamilton-Jacobi equation and Lie-
Poisson integrator. Computers Math. Applic., 30:67–74, (1995).
[LQHD07] X.S. Liu, Y.Y. Qi, J. F. He, and P. Z. Ding: Recent progress in symplectic algorithms
for use in quantum systems. Communications in Computational Physics, 2(1):1–53, (2007).
[LSD02a] X. S. Liu, L. W. Su, and P. Z. Ding: Symplectic algorithm for use in computing the
time independent Schrodinger equation. Int. J. Quant. Chem., 87:1–11, (2002).
[LSD02b] X.S. Liu, L.W. Su, and P. Z. Ding: Symplectic algorithm for use in computing the
time independent schrödinger equation. Int. J. Quant. Chem., 87(1):1–11, (2002).
[LYC06] H. Liu, J.H. Yuan, J.B. Chen, H. Shou, and Y.M. Li: Theory of large-step depth
extrapolation. Chinese Journal Geophys., 49(6):1779–1793, (2006).
[LZL93] X. Liao, Z. Zhao, and L. Liu: Application of symplectic algorithms in computation
of LCN. Acta Astronomica Sinica, 34(2):201–207, (1993).
[Men84] C.R. Menyuk: Some properties of the discrete Hamiltonian method. Physica D,
11:109–129, (1984).
[MF81] V.P. Maslov and M. V. Fedoriuk: Semi-classical approximation in quantum mechanics.
D. Reidel Publishing Company, Dordrecht Holland, First edition, (1981).
[MK95] H. Munthe-Kaas. Lie–Butcher theory for Runge–Kutta methods. BIT, 35(4):572–587,
(1995).
[MK98] H. Munthe-Kaas: Runge–Kutta methods on Lie groups. BIT, 38(1):92–111, (1998).
[MK99] H. Munthe-Kaas: High order Runge–Kutta methods on manifolds. Appl. Numer.
Math., 29:115–127, (1999).
[MKO99] H. Munthe-Kaas and B. Owren: Computations in a free Lie algebra. Phil. Trans.
Royal Soc. A, 357:957–981, (1999).
[MKQZ01] H. Munthe-Kaas, G. R. W. Quispel, and A. Zanna: Generalized polar decom-
positions on Lie groups with involutive automorphisms. Foundations of Computational
Mathematics, 1(3):297–324, (2001).
[MKZ97] H. Munthe-Kaas and A. Zanna: Numerical integration of differential equations on
homogeneous manifolds. In F. Cucker and M. Shub, editors, Foundations of Computational
Mathematics, pages 305–315. Springer Verlag, (1997).
[MM05] K.W. Morton and D.F. Mayers: Numerical Solution of Partial Differential Equations:
an introduction. Cambridge University Press, Cambridge, Second edition, (2005).
[MR99] J. E. Marsden and T. S. Ratiu: Introduction to Mechanics and Symmetry. Number 17
in Texts in Applied Mathematics. Springer-Verlag, second edition, (1999).
[MNSS91] R. Mrugała, J.D. Nulton, J.C. Schon, and P. Salamon: Contact structure in thermo-
dynamic theory. Reports on Mathematical Physics, 29:109C121, (1991).
[Mos62] J. Moser: On invariant curves of area-preserving mappings of an annulus. Nachr.
Akad. Wiss. Gottingen, II. Math.-Phys., pages 1–20, (1962).
[QC00] M. Z. Qin and J.B. Chen: Maslov asymptotic theory and symplectic algorithm. Chi-
nese Journal Geophys., 43(4):522–533, (2000).
[Qin89] M. Z. Qin: Cononical difference scheme for the Hamiltonian equation. Mathematical
Methodsand in the Applied Sciences, 11:543–557, (1989).
38 Bibliography
[Qin97a] M. Z. Qin: A symplectic schemes for the pde’s. AMS/IP studies in Advanced Math-
emateics, 5:349–354, (1997).
[QT90] G. D. Quinlan and S. Tremaine: Symmetric multistep methods for the numerical inte-
gration of planetary orbits. Astron. J., 100:1694–1700, (1990).
[QZ90] M. Z. Qin and M. Q. Zhang: Explicit Runge–Kutta–like Schemes to Solve Certain
Quantum Operator Equations of Motion. J. Stat. Phys., 60(5/6):839–843, (1990).
[QZ92] M. Z. Qin and W.J. Zhu: Construction of Higher Order Symplectic Schemes by Com-
position. Computing, 47:309–321, (1992).
[QZ93a] M. Z. Qin and W. J. Zhu: Volume-preserving schemes and numerical experiments.
Computers Math. Applic., 26:33–42, (1993).
[QZ93b] M. Z. Qin and W. J. Zhu: Volume-preserving schemes and applications. Chaos,
Soliton & Fractals, 3(6):637–649, (1993).
[QZ94] M. Z. Qin and W. J. Zhu: Multiplicative extrapolation method for constructing higher
order schemes for ode’s. J. Comput. Math., 12:352–356, (1994).
[Rut83] R. Ruth: A canonical integration technique. IEEE Trans. Nucl. Sci., 30:26–69, (1983).
[Sch44] E. Schrödinger: Scripta mathematica, 10:92–94, (1944).
[Shu93] H.B. Shu: A new approach to generating functions for contact systems. Computers
Math. Applic., 25:101–106, (1993).
[ST92a] P. Saha and S. Tremaine: Symplectic integrators for solar system dynamics. Astron.
J., 104:1633–1640, (1992).
[Syn44] J.L. Synge: Scripta mathematica, 10:13–24, (1944).
[Vog56] R. de Vogelaere: Methods of integration which preserve the contact transformation
property of the Hamiltonian equations. Report No. 4, Dept. Math., Univ. of Notre Dame,
Notre Dame, Ind., Second edition, (1956).
[War83] F. W. Warner: Foundations of Differentiable Manifolds and Lie Groups. GTM 94.
Springer-Verlag, Berlin, (1983).
[Wei77] A. Weinstein: Lectures on symplectic manifolds. In CBMS Regional Conference, 29.
American Mathematical Society,Providence,RI, (1977).
[wey39] H. weyl: The Classical Groups. Princeton Univ. Press, Princeton, Second edition,
(1939).
[WH91] J. Wisdom and M. Holman: Symplectic maps for the N -body problem. Astron. J.,
102:1528–1538, (1991).
[WHT96] J. Wisdom, M. Holman, and J. Touma: Symplectic Correctors. In Jerrold E. Mars-
den, George W. Patrick, and William F. Shadwick, editors, Integration Algorithms and
Classical Mechanics, volume 10 of Fields Institute Communications, pages 217–244.
Fields Institute, American Mathematical Society, July (1996).
[WJX01] B. Wang, Z. Ji, and Q. Xiao: the atmospheric dynamics of the equation Hamiltonian
algorithm. Chinese Journal Computational Physics, 18(1):289–297, (2001).
[WZJ95] B. Wang, Q. Zhen, and Z. Ji: The system of square conservation and Hamiltonian
systems. Science in China (series A), 25(7):765–770, (19950.
[Yos90] H. Yoshida: Construction of higher order symplectic integrators. Physics Letters A,
150:262–268, (1990).
[ZL93] Z. Zhao and L. Liu: The stable regions of triangular libration points of the planets .
Acta Astronomica Sinica, 34(1):56–65, (1993).
[ZL94] Z. Zhao and L. Liu: The stable regions of triangular libration points of the planets II.
Acta Astronomica Sinica, 35(1):76–83, (1994).
[ZLL92] Z. Zhao, X. Liao, and L. Liu: Application of symplectic integrators to dynamical
astronomy. Acta Astronomica Sinica, 33(1):33–41, (1992).
Chapter 1.
Preliminaries of Differentiable Manifolds
Before introducing the concept of differentiable manifold, we first explain what map-
ping is. Given two sets X, Y, and a corresponding principle, if for any x ∈ X, there
exists y = f (x) ∈ Y to be its correspondence, then f is a mapping of the set X into
the set Y , which is denoted as f : X → Y. X is said to be the domain of definition of
f , and f (x) = {f (x) | x ∈ X} ⊂ Y is said to be the image of f . If f (X) = Y , then
f is said to be surjective or onto; if f (x) = f (x ) ⇒ x = x , then f is said to be injec-
tive (one-to-one); if f is both surjective and injective (i.e., X and Y have a one-to-one
correspondence under f ), f is said to be bijective. For a bijective mapping f , if we
define x = f −1 (y), then f −1 : Y → X is said to be the inverse mapping of f . In ab-
stract algebra, a homomorphism is a structure-preserving map between two algebraic
structures (such as groups, rings, or vector spaces). For example, for two groups G and
G and a mapping f : G → G , a → f (a), if f (a, b) = f (a) · f (b), ∀a, b ∈ G, then f
is said to be a homomorphism from G to G . A homomorphism is a map from one al-
gebraic structure to another of the same type that preserves all the relevant structures,
i.e., properties such as identity element, inverse element, and binary operations. An
isomorphism is a bijective homomorphism. If f is a G → G homomorphic mapping,
and also a one-to-one mapping from G to G , then f is said to be a G → G isomor-
phic mapping. An epimorphism is a surjective homomorphism. Given two topological
spaces (x, τ ) and (y, τ ), if the mapping f : X → Y is one-to-one, and both f and its
inverse mapping f −1 : Y → X are continuous, then f is said to be a homeomorphism.
If f and f −1 are also differentiable, then the mapping is said to be diffeomorphism.
A monomorphism (sometimes called an extension) is an injective homomorphism. A
homomorphism from an object to itself is said to be an endomorphism. An endomor-
phism that is also an isomorphism is said to be an automorphism. Given two mani-
folds M and N , a bijective mapping f from M to N is called a diffeomorphism if
both f : M → N and its inverse f −1 : N → M are differentiable (if these functions
are r times continuously differentiable, f is said to be a C r -diffeomorphism).
Many differential mathematical methods and concepts are used in classical me-
chanics and modern physics: differential equations, phase flow, smooth mapping,
manifold, Lie group and Lie algebra, and symplectic geometry. If one would like to
construct a new numerical method, one needs to understand these basic theories and
concepts. In this book, we briefly explain manifold, symplectic algebra, and symplec-
tic geometry. In a series of books[AM78,Che53,Arn89,LM87,Ber00,Wes81] can be found these
materials.
40 1. Preliminaries of Differentiable Manifolds
ϕi (u1 , u2 ) = u2 , i = 1, 2; ϕi (u1 , u2 ) = u1 , i = 3, 4.
Xp : C ∞ −→ R,
(Xp + Yp )(f ) = Xp (f ) + Yp (f ),
(kXp )(f ) = kXp (f ), ∀ f ∈ C ∞ (p).
44 1. Preliminaries of Differentiable Manifolds
It is easy to verify that Tp M becomes the vector space that contains the above op-
eration, which is called the tangent space at the point p of the differential manifold
M.
Remark 1.21. By definition of the tangent vector, it is easy to know that if f is the
constant function, Xp (f ) = 0 for Xp ∈ Tp M .
Lemma 1.22. Let (U, ϕ) be the chart that contains p ∈ M , and let x1 , · · · , xm , ϕ(p)
= (a1 , · · · , am ) be the coordinate functions. If f ∈ C ∞ (p), then there exists a function
gi in some neighborhood W of p ∈ M , such that
m
f (q) = f (p) + (xi (q) − ai )gi (q), ∀ q ∈ W,
i=1
∂f ∂f ∂ ∂f ◦ ϕ−1
and gi (p) = i
where i = i (f ) = .
∂x p ∂x p ∂x p ∂ui ϕ(p)
1 ∂f ◦ ϕ−1
where g i (u) = 0
(su1 , · · · , sum ) d s (i = 1, · · · , m). Let g i (ϕ(q)) =
∂ui
gi (q), then gi is smooth on W , and satisfies
m
f (q) = f (p) + xi (q)gi (q),
i=1
∂f ◦ ϕ−1 ∂f
gi (p) = g i (O) = i
= i .
∂u O ∂x p
m
∂
Xp = Xp (xi ) .
∂xi p
i=1
1.1 Differentiable Manifolds 45
m
f = f (p) + (xi − ai )gi ,
i=1
then
∂f ∂
m m m
Xp (f ) = Xp [(x − a )gi ] =
i i
Xp (x ) i = i
Xp (xi ) i (f ).
i=1 i=1
∂x p i=1 ∂x p
Remark 1.24. By Theorem 1.23 we know: if the coordinates of Xp w.r.t. chart (U, ϕ)
are defined as (Xp (x1 ), · · ·, Xp (xm )), then Tp M and Rm are isomorphisms,
and the
∂
basis for Tp M corresponds exactly to the standard basis for R , i.e.,
m
→ ei =
∂xi p
(0, · · · , 1, 0, · · · , 0).
where fj = y j ◦ f .
46 1. Preliminaries of Differentiable Manifolds
Proof. Since
∂ ∂ ∂fk
f∗p (y k ) = i (y k ◦ f ) = i ,
∂xi p ∂x p ∂x p
n
∂fj ∂
= (y k ).
i ∂x p ∂y j f (p)
i,j=1
1.1.3 Submanifolds
The extension of the curve and surface on Euclidean space to the differentiable mani-
fold is the submanifold. In the following section, we focus on the definitions of three
submanifolds and their relationship. First, we describe a theorem.
1.1 Differentiable Manifolds 47
Example 1.32. Suppose f : R → S 1 , defined by f (t) = (cos t, sin t). Using the
chart of Example 1.11, we obtain
π π
cos t, t ∈ kπ − , kπ + ,
f (t) =
2 2
− sin t, t ∈ (kπ, (k + 1)π).
It is easy to prove that g(u,!0) = f(u) maps origin 0 to itself in Rn and the rank of
0
Jg (O ) = Jf(O) is n, where 0 denotes a m × (n − m) zero matrix, and by
In−m
the inverse function theorem, g is a diffeomorphism from a neighborhood of origin of
Rm to a neighborhood of origin of Rn . Shrink U1 , V1 so that they become U, V , and
−1
let ϕ = ϕ1 |U, ψ = g −1 ◦ (ψ|V ). Since ψ ◦ f ◦ ϕ−1 = g −1 ◦ ψ1 ◦ f ◦ ϕ−1
1 =g ◦f =
g(u, 0), the proposition is proved.
From the following example, we can see that the manifold topology of an immersed
submanifold may be inconsistent with its subspace topology and can be very complex.
1.1 Differentiable Manifolds 49
ϕ "−1
$β ◦ ϕ 1 −1 1
α (u , · · · , u ) = π ◦ ϕβ ◦ ϕα (u , · · · , u , 0, · · · , 0).
k k
It is easy to prove that G(gr(f)) = {(u, O ) | u ∈ ϕ(U )}, and the rank of
% &
Im O
JG (O, O ) =
−Df(O) In
" ),
W = (ϕ × ψ)−1 (U χ = G ◦ (ϕ × ψ)|W.
4. Embedded submanifolds
Definition 1.45. Let f : M → N be an injective immersion. If f : M → f (M ) is
a homeomorphism, where f (M ) has the subspace topology of N , then f (M ) is an
embedded submanifold of N .
f(u1 , · · · , um ) = (u1 , · · · , un ).
Proof. Take charts (U1 , ϕ1 ), (V, ψ1 ), p ∈ U1 , f (p) ∈ V, ϕ1 (p) = O ∈ Rm , ψ1 (f (p))
∂ fi
= O ∈ Rn and f (U1 ) ⊂ V . Since f is a submersion, Jf(O) = has rank
∂uj O
n, where f = (f1 , · · · , fn ). We assume that the first n rows of Jf(O) are linearly
independent. Let g : ϕ1 (U1 ) → ψ(V ) × Rm−n satisfy
from Rm → Rn .
Remark 1.53. By Definition 1.51, if f : M → N is a submersion at p ∈ M , then f
is a submersion in some neighborhood of p.
Remark 1.54. If f : M → N is a submersion, then f is an open mapping (i.e., open
set mapping to an open set). Furthermore, f (M ) is an open subset of N .
Let G be an open subset of M , ∀ q ∈ f (G). There exists a p ∈ G, s.t. f (p) = q.
Since f is a submersion, there exist charts (U, ϕ), (V, ψ), p ∈ U, q ∈ V , s.t. U ⊂ G,
and f : ϕ(U ) → ψ(V ), f(u1 , · · · , um ) = (u1 , · · · , un ). Let H = β(ϕ(U )), where
β(u1 , · · · , um ) = (u1 , · · · , un ), s.t. H ⊂ ψ(V ). Thus, ψ −1 (H) is a neighborhood of
q ∈ N , ψ −1 (H) ⊂ f (G), i.e., f (G) is an open subset of N .
Next, we consider under what condition would f −1 (q0 ) be a regular submanifold
of M , and ∀ q0 ∈ N be fixed.
Definition 1.55. Given f : M → N is smooth, p ∈ M , if f∗p : Tp M → Tf (p) N is
surjective, then p is said to be a regular point of f (i.e., f submerses at p), otherwise p
is said to be a critical point of f , and q ∈ N is called a regular value of f , if q ∈
/ f (M )
or q ∈ f (M ), but each p ∈ f −1 (q) is a regular point of f ; otherwise, q is called a
critical value of f .
Remark 1.56. When dim M < dim N , as a result of dim Tp M = dim M <
dim N = dim Tf (p) N , for q ∈ f (M ), p ∈ f −1 (q), p cannot be a regular point
of f . Hence, q ∈ N is a regular value of f ⇔ q ∈
/ f (M ).
Theorem 1.57. Let f : M → N be smooth, q ∈ N ; if q is a regular value of f ,
and f −1 (q) = Ø, then f −1 (q) is an (m − n)-dimensional regular submanifold of M .
Moreover, ∀p ∈ f −1 (q),
Tp {f −1 (q)} = ker f∗p .
1.1 Differentiable Manifolds 53
n+1
Example 1.59. Let f : Rn+1 → R, and f (u1 , · · · , un+1 ) = (ui )2 .
i=1
From the Jacobian matrix of f at (u1 , · · · , un+1 ), we know f is not a submersion
at (u1 , · · · , un+1 ) ⇔ u1 = · · · = un+1 = 0. Therefore, any non-zero real number is
a regular value of f . According to the Theorem 1.57, the n-dimensional unit sphere
S n = f −1 (1) is an n-dimensional regular submanifold on Rn+1 .
Example 1.60. Let f : R3 → R, and f (u1 , u2 , u3 ) = (a − (u1 )2 + (u2 )2 )2 +
(u3 )2 , a > 0.
The assumption tells us that any non-zero real number is a regular point of f . Then,
0 < b2 < a2 is a regular value of f . Therefore, by Theorem 1.57, T 2 = f −1 (b2 ) is a
2-dimensional regular submanifold on R2 .
If M is a regular submanifold of M , then dim M − dim M =codim M is called
the M -codimension of M . Denote M = {p ∈ M | fi (p) = 0 (i = 1, · · · , k)} and
consider the mapping
i∗p (Tp Z ) + Tp Z = Tp N,
i.e.,
1.1 Differentiable Manifolds 55
Tp Z + Tp Z = Tp N.
We assume that f : M → N is smooth, and Z is a k-codimensional regular
submanifold of N , p ∈ M, f (p) = q ∈ Z. According to the Proposition 1.61, there
exists a submanifold chart (V, ψ) of N that contains q, s.t. π ◦ ψ : V → Rk is a
submersion, and Z ∩ V = (π ◦ ψ)−1 (O). Now, take a neighborhood of p in M , s.t.,
f (U ) ⊂ V , then π ◦ ψ ◦ f : U → Rk .
Proposition 1.65. f p Z ⇔ π ◦ ψ ◦ f : U → Rk submerses at p.
Tp {f −1 (Z)} = f∗p
−1
{Tf (p) Z}.
−1
= f∗p {Tf (p) Z}.
ϕU (Xp ) = (ϕ(p); a1 , · · · , am ),
obviously ϕU is a 1 to 1 mapping.
Note that as (U, ϕ) takes all the charts on M , all the corresponding π −1 (U ) con-
stitutes a covering of T M . Hence, if the topology of π −1 (U ) is given, the subset of
π −1 (U ) is open, iff the image of ϕU is an open set of ϕ(U ) × Rm . It is easy to prove
that by the 1 to 1 correspondence of ϕU , the topology of ϕU on the Rm ×Rm = R2m
subspaces can be lifted on π −1 (U ). The topology on T M can be defined as follows:
W is called an open subset of T M , iff W ∩ π −1 (U ) is an open subset of π −1 (U ). It
is easy to deduce that T M constitutes a topological space that satisfies the following
conditions:
1◦ T M is a Hausdorff space that has countable bases.
2◦ π −1 (U ) is an open subset of T M , and ϕU is a homeomorphism from π −1 (U )
to an open subset of R2m .
Furthermore, it can be proved that the manifold structure on T M can be naturally
induced from the manifold structure on M . We say that {(π −1 (U ), ϕU )} = A is a
smooth atlas of T M . For any chart (π −1 (U ), ϕU ), there exists a (π −1 (V ), ψV ) ∈
1.2 Tangent Bundle 57
(u; a1 , · · · , am ) = u.
π
By the definition of submersion, π is a submersion.
Given below are examples of two trivial tangent bundles (if there exists a diffeo-
morphism from its tangent bundle T M to M × Rm , and this diffeomorphism limited
on each fiber of T M (Tp M ) is a linear isomorphism from Tp M to {p} × Rm ).
Example 2.3. Let U be an open subset of Rm and T U U × Rm .
∂
∂ (i = 1, · · · , m) is the basis of
∀ Xu ∈ T U, Xu = ai i , where
i
∂u u ∂ui u
Tu U . Then, it is easy to prove that
Xu −→ (u; a1 , · · · , am )
is a diffeomorphism from T U to U × Rm . Moreover, since each fiber Tu U of T U is
a linear space, maps limited on Tu U is a linear isomorphism from Tu U to {u} × Rm ,
i.e., T U is a trivial tangent bundle.
where x, y are the coordinate functions on (U, ϕ), (V, ψ) respectively. When p ∈ U ∩
V , we have
∂ ∂y ∂ ∂
= = .
∂x p ∂x p ∂y p ∂y p
Therefore, f has the definition and is a 1 to 1 correspondence. Moreover, f and f −1
are smooth. Hence, T S 1 is a trivial tangent bundle.
Apart from trivial tangent bundles, there exists a broad class of nontrivial tangent
bundles. For an example, T S 2 is a nontrivial tangent bundle.
T f |Tp M = f∗p , ∀ p ∈ M,
ψV ◦ T f ◦ ϕ−1 1
U (u; a , · · · , a )
m
∂f1 ∂fn
= ψ ◦ f ◦ ϕ−1
U ; ai i ,···, ai i ,
i
∂x ϕ−1 (u) i
∂x ϕ−1 (u)
T (g ◦ f ) = T g ◦ T f.
2. Orientation
Next, we introduce the concept of orientation for differentiable manifolds.
Given V as a m-dimensional vector space, {e1 , · · · , em }, {e1 , · · · , em } as V ’s two
m
ordered bases, if ej = aij ei (j = 1, · · · , m), then
i=1
where A = (aij )m×m . If det A > 0, we call {ei } and {ej } concurrent; otherwise,
if det A < 0, we call {ei } and {ej } reverse. Then, a direction μ of V can be ex-
pressed by a concurrent class [{ej }] equivalent to {ej }. The other direction −μ can
be expressed by an equivalent class to the reverse direction of {ej }. (V, μ) is called an
orientable vector space.
Let (V, μ), (W, ν) be two orientable vector spaces. A : V → W is a linear isomor-
phism from V to W . If the orientation of W , which is induced by A, is consistent with
ν, i.e., Aμ = ν, then A preserves orientations. Otherwise, A reverses orientations. In
the below section, we extend the orientation concept to differentiable manifolds.
are all linear isomorphisms that preserves orientations, where (U, ϕ) is a chart that
contains p, and
∂ ∂
νϕ(q) = , · · · , .
∂u1 ϕ(q) ∂um ϕ(q)
Then, μ = {μp | p ∈ M } is the orientation on M , and (M, μ) is called an orientable
differentiable manifold.
Remark 2.10. The Definition 2.9 shows that if (M, μ) is an orientable differentiable
manifold, W is an open subset of M , then ∀p ∈ M and there exists an orientation μp
of Tp M . This gives an orientation on W , denoted by μ|W . Then, (W, μ|W ) is also an
orientable differentiable manifold. Specifically, if (U, ϕ) is a chart on M , then (U, μp )
is an orientable differentiable manifold.
Remark 2.10 shows that M may be locally orientable. Next, we discuss how to
construct a global orientation.
Sufficiency. Let A be an atlas that satisfies all the properties of the proposition.
Choose (Uα , ϕα ), (Uβ , ϕβ ) ∈ A and Uα ∩ Uβ = Ø, and use x1 , · · · , xm and
y 1 , · · · , y m to represent the coordinate functions of (Uα , ϕα ), (Uβ , ϕβ ) respectively.
Note that
∂ ∂ ∂ ∂
, · · · , = , · · · , J −1 (ϕα (q)),
∂x1 p ∂xm p ∂y 1 p ∂y m p ϕβ ◦ϕα
i.e., M is orientable.
Remark 2.14. By the Proposition 2.13, any connected open set on an orientable dif-
ferentiable manifold M has two and only two orientations.
Remark 2.15. Let (U, ϕ), (V, ψ) be two charts on M , and U and V be connected. If
U ∩ V = Ø, then det Jψ◦ϕ−1 preserves the orientation on ϕ(U ∩ V ).
ϕ± : U± → R, s.t.
u1 −u1
ϕ+ (u1 , u2 ) = , ϕ− (u1 , u2 ) = .
1 + u2 u2 − 1
Since
1
ϕ+ ◦ ϕ−1
− (u) = − , ∀ u ∈ ϕ− (U+ ∩ U− ),
u
we have
1
det Jϕ+ ◦ϕ−1 (u) = > 0, ∀ u ∈ ϕ− (U+ ∩ U− ).
− u2
Similarly
det Jϕ− ◦ϕ−1 (u) > 0, ∀ u ∈ ϕ+ (U+ ∩ U− ),
+
i.e., S 1 is orientable.
Example 2.17. Möbius strip is a non-orientable surface. Define equivalent relation“∼”
on [0, 1] × (0, 1):
[0, 1] × (0, 1)\ ∼ is a Möbius strip, A = {(U, ϕ), (V, ψ)} is its smooth atlas
1
U = M \{0} × (0, 1), V = M\ × (0, 1),
2
1 1
ϕ : U −→ (0, 1) × (0, 1), ψ : V −→ − , × (0, 1),
2 2
which satisfies:
ϕ(u, v) = (u, v),
⎧
⎪ 1
⎨ (u, v), 0≤ ,
2
ψ(u, v) =
⎪
⎩ (u − 1, 1 − v), 1
< u ≤ 1,
2
⎧
⎪ 1
⎨ (u, v), (u, v) ∈ 0, × (0, 1),
−1 2
ψ ◦ ϕ (u, v) =
⎪
⎩ (u − 1, 1 − v), (u, v) ∈ 1 , 1 × (0, 1),
2
i.e., ⎧
⎪ 1
(u, v) ∈ 0,
⎨ 1, × (0, 1),
2
det Jψ◦ϕ−1 (u, v) =
⎪
⎩ −1, (u, v) ∈
1
, 1 × (0, 1).
2
By the Remark 2.15, Möbius strip is a nonorientable surface.
62 1. Preliminaries of Differentiable Manifolds
by
= ϕU ◦ X ◦ ϕ−1 : ϕ(U ) −→ ϕ(U ) × Rm ,
X
X(u) = (u; a1 ◦ ϕ−1 (u), · · · , am ◦ ϕ−1 (u)),
we know, if X is smooth, then a1 , · · · , am are smooth too. Since
∂f
(Xf )(p) = Xp f = ai (p) i , ∀ p ∈ U,
i
∂x p
∂
Xq = xi )
Xq (" , ∀ q ∈ U,
i
∂xi q
π2 ◦ (f∗ X) = π2 ◦ (T f ◦ X ◦ f −1 )
= f ◦ (π1 ◦ X) ◦ f −1 = I.
Remark 2.24. Let X be a smooth vector field on the differentiable manifold M , and
(U, ϕ) be a chart on M . By Proposition 2.23, we have ϕ∗ (X | U ) to be a smooth
vector field of ϕ(U ).
then
∂ m
∂(ϕ ◦ c)i ∂
(ϕ ◦ c)∗t = , ∀ t ∈ J,
∂t t ∂t t ∂ui ϕ◦c(t)
i=1
⎧ i
⎨ d u = ai (u1 , · · · , um ), i = 1, · · · , m,
dt
⎩
u(0) = ϕ(p).
∀ x ∈ Rn , A(x) ∈ Tx Rn .
Since (e1 )x , · · · , (en )x form a basis on Tx Rn , we can write ∂ , · · · , ∂ , there-
∂x1 ∂xn
fore
n
A(x) = Ai (x)(ei )x .
i=1
If Ai (x) ∈ C ∞ , then A(x) is called a smooth vector field on Rn . The set of all smooth
vector fields on Rn is denoted by X (Rn ). For any vector A(x), B(x) ∈ X (Rn ),
define:
(αA + βB)(x) = αA(x) + βB(x), α, β ∈ R,
n
A(x) = Ai (x)(ei )x = (A1 (x), · · · , An (x)) ,
i=1
( )
V = x ∈ R3 | x = α1 a1 + α2 a2 + α3 a3 , 0 ≤ α1 , α2 , α3 ≤ 1
a spanned parallelepiped by vectors a1 , a2 , a3 . We introduce a new operation, ∧ be-
tween a1 , a2 , a3 as follows:
a11 a12 a13
a1 ∧ a2 ∧ a3 = a21 a22 a23 .
a31 a32 a33
The geometric meaning of a1 ∧a2 ∧a3 is the orientable volume of V , where orientation
means the sign of the volume is positive or negative. If the right hand law is followed,
the volume has the plus sign, otherwise it has the minus sign . It is easy to see that
operation ∧ satisfies the following laws:
1◦ Multilinear. Let a2 = βb + γc, b, c be vectors, β, γ be real numbers. Then,
a1 ∧ (βb + γc) ∧ a3 = β(a1 ∧ b ∧ a3 ) + γ(a1 ∧ c ∧ a3 ).
2◦ Anti-commute
a1 ∧ a2 ∧ a3 = −a2 ∧ a1 ∧ a3 ,
a1 ∧ a2 ∧ a3 = −a3 ∧ a2 ∧ a1 ,
a1 ∧ a2 ∧ a3 = −a1 ∧ a3 ∧ a2 .
From 2◦ we know that if a1 , a2 , a3 has two identical vectors, then a1 ∧a2 ∧a3 = 0.
Example 3.1. Let e1 , e2 , e3 be a basis in R3 , which are not necessarily orthogonal,
and let a1 , a2 , a3 be three vectors in R3 , which can be represented by
a1 = a11 e1 + a12 e2 + a13 e3 ,
a2 = a21 e1 + a22 e2 + a23 e3 ,
a3 = a31 e1 + a32 e2 + a33 e3 .
By multilinearity and anti-commutativity of ∧, after the computation, we have
a11 a12 a13
a1 ∧ a2 ∧ a3 = a21 a22 a23 e1 ∧ e2 ∧ e3 .
a31 a32 a33
and so
ω = ω(e1 )x1 + ω(e2 )x2 + · · · + ω(en )xn .
Thus, x1 , · · · , xn is a basis on Λ1 (Rn ), Λ1 (Rn ) = {xi }i=1,···,n .
Example 3.4. If F is a uniform force field on a Euclidean space R3 , then its work A
on a displacement ξ is a 1-form acting on ξ,
ωF (ξ) = (F, ξ) = F1 a1 + F2 a2 + F3 a3 , ξ = a1 e1 + a2 e2 + a3 e3
or
ωF = F1 x1 + F2 x2 + F3 x3 .
2. 2-Forms
Definition 3.5. An exterior form of degree 2 (or a 2-form) is a bilinear, skew-
symmetric function ω 2 : Rn × Rn → R, i.e.,
0 = ω 2 (ξ1 + ξ2 , ξ1 + ξ2 )
= −ω 2 (ξ1 , ξ1 ) + ω 2 (ξ1 , ξ2 ) + ω 2 (ξ2 , ξ1 ) + ω 2 (ξ2 , ξ2 )
= ω 2 (ξ1 , ξ2 ) + ω 2 (ξ2 , ξ1 ).
Example 3.7. Let S(ξ1 , ξ2 ) be the oriented area of the parallelogram constructed on
the vector ξ1 and ξ2 of the oriented Euclidean plane R2 , i.e.,
ξ11 ξ12
S(ξ1 , ξ2 ) = ,
ξ21 ξ22
where
ξ1 = ξ11 e1 + ξ12 e2 , ξ2 = ξ21 e1 + ξ22 e2 .
Example 3.8. Let v be a given vector on the oriented Euclidean space R3 . The triple
scalar product on other two vectors ξ1 and ξ2 is a 2-form:
v1 v2 v3
ω(ξ1 , ξ2 ) = (v, [ξ1 , ξ2 ]) = ξ11 ξ12 ξ13 ,
ξ21 ξ22 ξ23
3
3
where v = vi eji , ξj = ξji ei (j = 1, 2).
i=1 i=1
3. k-Forms
We denote the set of all permutations of the set {1, 2, · · · , k} by Sk and its element by
where σ = (i1 , i2 , · · · , ik ) ∈ Sk .
The set of all k-forms in Rn is denoted by Λk (Rn ). It forms a real vector space if
we introduce operations of addition.
it
k
Question 3.11. Show that if ηj = aji ξi (j = 1, · · · , k), then
i=1
Definition 3.12. For ω1 and ω2 ∈ Λ1 (Rn ), the exterior product of ω1 and ω2 denoted
by ω1 ∧ ω2 is defined by the formula
ω1 (ξ1 ) ω2 (ξ1 )
(ω1 ∧ ω2 )(ξ1 , ξ2 ) = , ξ1 , ξ2 ∈ Rn ,
ω1 (ξ2 ) ω2 (ξ2 )
1.3 Exterior Product 69
which denotes the oriented area of the image of the parallelogram with sides ω(ξ1 )
and ω(ξ2 ) on the ω1 , ω2 plane.
It is not hard to verify that ω1 ∧ ω2 really is a 2-form and has properties
ω1 ∧ ω2 = −ω2 ∧ ω1 ,
(λ1 ω1 + λ2 ω1 ) ∧ ω2 = λ1 ω1 ∧ ω2 + λ2 ω1 ∧ ω2 .
Now suppose we have chosen a system of linear coordinates on Rn , i.e., we are given
n independent 1-forms, x1 , x2 , · · · , xn . We will call these forms basic. The exterior
products of the basic forms are the 2-forms xi ∧ xj . By skew-symmetry,
xi ∈ Λ1 (Rn ), xi ∧ xi = 0,
xi ∧ xj = −xj ∧ xi ,
xi (ξ1 ) xj (ξ1 )
(xi ∧ xj )(ξ1 , ξ2 ) =
xi (ξ2 ) xj (ξ2 )
ai aj
= = ai bj − aj bi ,
bi bj
where ξ1 = ai ei , ξ2 = bi ei . It is the oriented area of the parallelogram with
i i
sides (xi (ξ1 ), xi (ξ2 )) and (xj (ξ1 ), xj (ξ2 )) in the (xi , xj )-plane.
For any ω ∈ Λ2 (Rn ),
n
n
ω(ξ1 , ξ2 ) = ω(ai ei , bj ej ) = ai bj ω(ei , ej )
i,j=1 i,j=1
= (ai bj − aj bi )ω(ei , ej ) = ω(ei , ej )(xi ∧ xj )(ξ1 , ξ2 ),
i<j i<j
where ξ1 = ai ei , ξ2 = bi ei . Thus,
i i
ω= ω(ei , ej )xi ∧ xj ,
i<j
Thus, {xi ∧xj }i<j are linearly independent and they form a base of Λ2 (Rn ), which
n
implies that the dimension of Λ2 (Rn ) is .
2
70 1. Preliminaries of Differentiable Manifolds
ω = P x2 ∧ x3 + Qx3 ∧ x1 + Rx1 ∧ x2 .
2. Exterior monomials
Proof. Here we only prove 5◦ , the others are easy. By the linearity of the exterior
product,
* k + * k +
β1 ∧ · · · ∧ βk = a1i1 ωi1 ∧ · · · ∧ akik ωik
i1 =1 ik =1
k
= a1i1 · · · akik ωi1 ∧ · · · ∧ ωik
i1 ,···,ik =1
= a1i1 · · · akik ε(i1 , · · · , ik )ω1 ∧ · · · ∧ ωk (by 2◦ )
i1 ,···,ik ∈νk
Theorem 3.16. {xi1 ∧ · · · ∧ xik }i1 <···<ik form a basis on Λk (Rn ),and so the dimen-
n
sion of Λk (Rn ) = .
k
n
Proof. For ξi = ξij ej (i = 1, · · · , k), ξi ∈ Rn , then
j=1
n
n
ω(ξ1 , · · · , ξk ) = ω ξi,j1 ej1 , · · · , ξkjk ejk
j1 =1 jk =1
n
= ξij1 · · · ξkjk ω(ej1 , · · · , ejk )
j1 ,···,jk =1
= ε(j1 , · · · , jk )
i1 <···<ik (j1 ,···,jk )∈νk (i1 ,···,ik )
So
ω= ω(ei1 , · · · , eik )xi1 ∧ · · · ∧ xik .
i1 <···<ik
Thus, {xi1 ∧ · · · ∧ xik }i1 <···<ik generate Λk (Rn ). Obviously, they are linearly inde-
pendent. Consequently, they form a basis on Λk (Rn ).
In particular,
n
dim Λk (Rn ) = .
k
n
If k = n, then dim Λn (Rn ) = = 1, ∀ ω ∈ Λn (Rn ), and there must be
n
ω = ax1 ∧ · · · ∧ xk , for some number a ∈ R.
For k > n,
dim (Λk (Rn )) = 0, Λk (Rn ) = {0}.
Therefore, the theorem is completed.
2◦ ω1 ∧ ω2 = (−1)kl ω2 ∧ ω1 , ω1 ∈ Λk , ω2 ∈ Λl .
The proof is left to the reader. Now we turn to prove Proposition 3.22. For this we
only need to prove 4◦ of Lemma 3.23.
1.3 Exterior Product 73
Proof. By 4◦ in Lemma 3.23, we only need to prove that if ξi = ξi +1 , then (ω k ∧
ω l )(ξ1 , · · ·, ξi , ξi +1 , · · · , ξk+l ) = 0.
Consider the terms of the right hand side in (1.2). If i , i + 1 ∈ (i1 , · · · , ik ), then
ω k (ξi1 , · · · , ξik ) = 0. Therefore,
σ = (i1 , · · · , i , · · · , ik , j1 , · · · , i + 1, · · · , jl ),
σ = (i1 , · · · , i + 1, · · · , ik , j1 , · · · , i , · · · , jl ).
(ω k ∧ ω l )(ξ1 , · · · , ξk+l ) = 0.
we get skew-commutativity.
In order to prove associativity, we first prove that for monomials the exterior prod-
uct defined by Definition 3.19 coincides with the exterior product in Definition 3.17.
Since we have not get proved the equivalence of the Definition 3.17 of exterior product
of k 1-forms with the Definition 3.19 , we will temporarily denote the exterior product
of k 1-forms by the symbol ∧, so that our monomials have the form
Proof.
· det ωj (ξjm )
k+1≤j≤k+l; 1≤m≤k
Thus,
ω1 ∧ω2 = ω1 ∧ ω2 ,
ω1 ∧ω2 ∧ω3 = (ω1 ∧ω2 ) ∧ ω3 = (ω1 ∧ ω2 ) ∧ ω3 ,
ω1 ∧ω2 ∧ω3 = ω1 ∧ ω2 ∧ω3 = ω1 ∧ (ω2 ∧ ω3 ).
It follows that
(ω1 ∧ ω2 ) ∧ ω3 = ω1 ∧ (ω2 ∧ ω3 ),
and denoted by ω1 ∧ ω2 ∧ ω3 . Thus, ω1 ∧ω2 ∧ω3 = ω1 ∧ ω2 ∧ ω3 .
In general, we have
ωk = ai1 ···ik xi1 ∧ · · · ∧ xik ,
ωl = bj1 ···jl xj1 ∧ · · · ∧ xjl ,
ωm = ch1 ···hm xh1 ∧ · · · ∧ xhm .
= ω k ∧ (ω l ∧ ω m ).
Λ = Λ0 +̇Λ1 +̇ · · · +̇Λn .
ω = ω0 + ω1 + · · · + ωn , ωi ∈ Λ i ,
and this kind of expression is unique. In Λ there is not only algebraic structure of the
linear space, but also the definition of the exterior product. Direct sum Λ is the Grass-
mann algebra produced by the linear space which contained the entire real number
field and the linear space.
1, x1 , · · · , xn , xi ∧ xj (i < j), · · · , x1 ∧ x2 ∧ · · · ∧ xn
n
n
form the basis of Λ, whose dimension is dim(Λ) = = 2n .
i
i=0
There is no strict definition on how to define df for a smooth function and dx for
dx1 , · · · , dxn , in classical mathematical analysis. The differential of the independent
variable is equal to its increment in classical mathematical analysis, which is improper
in a general sense. Here, we always regard dx1 , · · · , dxn as some basis of a linear
space, which is called the differential space.
76 1. Preliminaries of Differentiable Manifolds
is called the tangent bundle on Rn , see Section 1.2. Notice that the tangent bundle
T (Rn ) consists of all fixed vectors on Rn .
The mapping π : T Rn → Rn defined by the following formula:
π(ξx ) = x, ∀ξx ∈ Tx Rn
is called the tangent bundle projection. π −1 (x) = Tx Rn is called the fiber of the
tangent bundle over the point. The dual space of Tx Rn denoted by Tx∗ Rn is called the
cotangent vector space to Rn at x consisting of all linear functions from Tx Rn into
Rn . Its element is called a covector (covariant vector) or a cotangent vector to Rn at
x, and ,
T ∗ (Rn ) = Tx∗ Rn
x∈Rn
Then, Ω∞ (Rn ) forms a C ∞ (Rn )-module, i.e., vector space over a ring. ∀ ω k ∈
Ωk (Rn ), ω l ∈ Ωl (Rn ), define their exterior product ω k ∧ ω l ∈ Ωk+l (Rn ) as
By the Theorem 3.24, the exterior product of differential forms defined above is dis-
tributive, skew-symmetric, and associative.
If f : Rn → R is differentiable, then Df (x) ∈ Λ1 (Rn ), where Df is the deriva-
tive of f (x) at x. Thus, we get a differential 1-form df ∈ Ω1 (Rn ), defined as
n
df (ξx ) = Df (x)ξ = Di f (x)ξi , ξ= ξi ei .
i=1
or
dxi ((ej )x ) = Dxi (ej ) = δij .
Thus, dx1 , · · · , dxn form the dual basis to (e1 )x , · · · , (en )x . ∀ ω ∈ Ωk (Rn ), ω can
be written as
ω(x) = ai1 ···ik (x)dxi1 ∧ · · · ∧ dxik ,
i1 <···<ik
∞
where ai1 ···ik (x) ∈ C (R ). n
df = D1 f dx1 + · · · + Dn f dxn
or in a classical notation
∂f ∂f
df = dx1 + · · · + n dxn ,
∂x1 ∂x
since
n
n
df (ξx ) = D f (x)(ξ) = ξi Di f = Di f dxi (ξx ), ∀ ξx ∈ Tx Rn .
i=1 i=1
Theorem 4.2. Every differential k-form on the space Rn with a given coordinate
system x1 , · · · , xn can be represented uniquely in the form
ωk = ai1 ···ik (x)dxi1 ∧ · · · ∧ dxik ,
i1 <···<ik
Theorem 4.3. Every differential 1-form on the space Rn with a given coordinate
system x1 , · · · , xn can be represented uniquely with smooth function ai (x) as follows:
ω = a1 (x)dx1 + · · · + an (x)dxn .
Example 4.4. Calculate the value of the forms ω = dr2 (r2 = (x1 )2 + (x2 )2 ) on the
vectors ξ1 , ξ2 , ξ3 . (Fig. 4.1), the results in full in the Table 4.1.
x2
6
2 ξ2 ξ3
1 = ~
ξ1
6
- x1
0 1 2 3
Fig. 4.1. Example 4.4 graphical representations
x2
6
3
6
η3
ξ3
2 η1
6
η2
1 - ξ1
ξ2
R - x1
0 1 2
Fig. 4.2. Example 4.5 graphical representations
Example 4.6. Calculate the value of the forms ω1 = dx2 ∧ dx3 , ω2 = x1 dx3 ∧
dx2 , ω3 = dx3 ∧ dr2 on the vectors ξ, η at the point x, where r2 = (x1 )2 + (x2 )2 +
(x3 )2 , ξ = (1, 1, 1) , η = (1, 2, 3) , x = (2, 0, 0).
The detailed calculation is shown bellow as follows:
dx2 (ξ) dx3 (ξ) 1 1
2 3
ω1 (ξ, η) = dx ∧ dx (ξ, η) = = = 1,
dx2 (η) dx3 (η) 2 3
ω2 (ξ, η) = 2 · dx3 ∧ dx2 (ξ, η) = −2ω1 (ξ, η) = −2,
ω3 (ξ, η) = dx3 ∧ dr2 (ξ, η)
= dx3 ∧ (2x1 dx1 + 2x2 dx2 + 2x3 dx3 )(ξ, η)
= 2x1 dx3 ∧ dx1 (ξ, η) − 2x2 dx2 ∧ dx3 (ξ, η)
= 2 · 2 · dx3 ∧ dx1 (ξ, η) − 2 · 0 · dx2 ∧ dx3 (ξ, η)
dx3 (ξ) dx1 (ξ) 1 1
= 4 = 4 = −8.
dx3 (η) dx1 (η) 3 1
80 1. Preliminaries of Differentiable Manifolds
(f ∗ ω)(x) = f ∗ ω(x).
Thus, we have
n
∂f i j
f ∗ (dy i ) = dx .
j=1
∂xj
we have
!
i1 · · · ik
f ∗ω = ai1 ,···,ik (f (x))Δ dxj1 ∧ · · · ∧ dxjk ,
j1 · · · jk
i1 <···<ik ; j1 <···<jk
! !
i1 · · · ik i1 · · · ik ∂y i
where Δ is the -minor of matrix i = 1, · · · , m .
j1 · · · jk j1 · · · jk ∂xj j = 1, · · · , n
Proof.
* +
∗ ∗
f ω =f ai1 ,···,ik (y)dy ∧ · · · ∧ dy
i1 ik
i1 <···<ik
= ai1 ,···,ik (f (x))f ∗ (dy i1 ) ∧ · · · ∧ f ∗ (dy ik )
i1 <···<ik
⎛ ⎞ ⎛ ⎞
n
∂y i1 n
∂y ik
= ai1 ,···,ik (f (x)) ⎝ j1
dxj1 ⎠ ∧ · · · ∧ ⎝ jk
dxjk ⎠
i1 <···<ik j =1
∂x j =1
∂x
1 k
n
∂y i1
∂y ik
= ai1 ,···,ik (f (x)) j1
· · · jk dxj1 ∧ · · · ∧ dxjk
i1 <···<ik j1 ,···,jk =1
∂x ∂x
n % & i1
j1 · · · jk ∂y
= ai1 ,···,ik (f (x)) ε
j1 · · · jk
∂xj1
i1 <···<ik j1 <···<jk (j1 ,···,jk )∈νk
ik
∂y
··· dx
j1
∧ · · · ∧ dxjk
∂xjk
% &
i1 · · · ik
= ai1 ,···,ik (f (x))Δ dxj1 ∧ · · · ∧ dxjk .
j1 · · · jk
i1 <···<ik ; j1 <···<jk
For proof of 5◦ :
Proof.
Therefore, (h ◦ f )∗ = f ∗ ◦ h∗ .
where k = 0, 1, · · · , n.
Exterior algebra may be represented by the local coordinate system of M as
ω= ai1 ,···,ik dxi1 ∧ · · · ∧ dxik ,
1≤i1 <···<ik ≤n
n
∂ai 1 ,···,ik
= dxj ∧ dxi1 ∧ · · · ∧ dxik .
1≤i1 <···<ik ≤n j=1
∂xj
Then,
ωk ∧ ωl = ai1 ,···,ik (x)bj1 ,···,jl (x)dxi1 ∧ · · · ∧ dxik ∧ dxj1 ∧ · · · ∧ dxjl .
i1 <···<ik ; j1 <···<jl
By definition,
n
∂(ab) i
d(ω k ∧ ω l ) = dx ∧ dxi1 ∧ · · · ∧ dxik ∧ dxj1 ∧ · · · ∧ dxjl
i1 <···<ik ; j1 <···<jl i=1
∂xi
* +
n
∂ai1 ···ik ∂bj1 ···jl
= bj1 ···jl i
+ ai1 ···ik dxi ∧ dxi1 ∧ · · · ∧ dxik
i1 <···<ik ; j1 <···<jl i=1
∂x ∂xi
∧dxj1 ∧ · · · ∧ dxjl
n
∂ai1 ···ik i
= bj1 ···jl dx ∧ dxi1 ∧ · · · ∧ dxik ∧ dxj1 ∧ · · · ∧ dxjl
i1 <···<ik ; j1 <···<jl i=1
∂xi
n
∂bj1 ···jl
+ ai1 ···ik (−1)k dxi ∧ dxi1 ∧ · · · ∧ dxik ∧ dxj1 ∧ · · · ∧ dxjl
i1 <···<ik ; j1 <···<jl i=1
∂xi
= dω k ∧ ω l + (−1)k ω k ∧ dω l .
n
∂2a
d(dω) = i j
dxi ∧ dxj ∧ dxi1 ∧ · · · ∧ dxik
i1 <···<ik i,j=1
∂x ∂x
% ∂2a ∂2a
&
= i j
− j i
dxi ∧ dxj ∧ dxi1 ∧ · · · ∧ dxik
i1 <···<ik i<j
∂x ∂x ∂x ∂x
= 0,
since
84 1. Preliminaries of Differentiable Manifolds
∂2a ∂2a
= .
∂xi ∂xj ∂xj ∂xi
4◦ By Theorem 4.8, 1 ◦ ∀ ω = g ∈ C ∞ (Rn ), we have:
*m +
∂g m
∂g
∗
f dg = f ∗
i
dy i
= i
◦ f · f ∗ (dy i )
i=1
∂y i=1
∂y
n m
∂g(f (x)) ∂f i j
= · j dx
j=1 i=1
∂y i ∂x
n
∂g(f (x))
= dxj = d(g ◦ f ) = df ∗ g.
j=1
∂xj
Furthermore, ∀ ω ∈ Ωk (Rm ),
* +
m
∂a i
∗ ∗
f (dω) = f i
dy ∧ dy i1 ∧ · · · ∧ dy ik
i1 <···<ik i=1
∂y
* +
∗
=f da ∧ dy ∧ · · · ∧ dy
i1 ik
i1 <···<ik
= f ∗ (da) ∧ f ∗ dy i1 ∧ · · · ∧ dy ik
i1 <···<ik
= d(f ∗ a) ∧ f ∗ dy i1 ∧ · · · ∧ dy ik
i1 <···<ik
* +
∗ ∗
=d f a∧f dy ∧ · · · ∧ dy
i1 ik
(by 3◦ )
i1 <···<ik
= df ∗ ω.
Therefore, the theorem is completed.
Proof. We will construct an R-linear mapping H : Ωk (A) → Ωk−1 (A), such that
d ◦ H + H ◦ d = id: Ωk (A) → Ωk (A), i.e., ω = d ◦ H(ω) + Hdω. Then, from dω = 0,
if follows that ω = d(H(ω)). Taking η = H(ω), we get
ω = dη.
Let
ω= ai1 ···ik (x)dxi1 ∧ · · · ∧ dxik .
i1 <···<ik
Define
k %- 1
j−1
H(ω)(x) = (−1) tk−1 ai1 ···ik (tx)dtxij
i1 <···<ik j=1 0
&
.
· dx ∧ · · · ∧ dx ∧ · · · ∧ dx
i1 i j ik
,
k %- 1 &
dH(ω) = (−1)j−1 tk−1 a(tx)dt dxij ∧ dxi1 ∧ · · · ∧ dxik
i1 <···<ik j=1 0
* n -
+
k
j−1
1
k ∂a(tx) i
+ (−1) t dx dt xij ∧ dxi1 ∧ · · · ∧ dxik
i1 <···<ik j=1 i=1 0 ∂xi
%- 1 &
= k tk−1 a(tx)dt dxi1 ∧ · · · ∧ dxij ∧ · · · ∧ dxik
i1 <···<ik 0
k n -
1
∂a(tx)
+ (−1)j−1 tk dtxij dxi ∧ dxi1 ∧ · · ·
i1 <···<ik j=1 i=1 0 ∂xi
.
∧dxij
∧ · · · ∧ dxik . (4.1)
n
∂a i
dω = i
dx ∧ dxi1 ∧ · · · ∧ d xik ,
i1 <···<ik i=1
∂x
' n - 1
∂a(tx) i i1
Hdω = tk i
dt x dx ∧ · · · ∧ dxik
i <···<i i=1 0 ∂x
1 k
n
k - 1
∂a(tx)
+ (−1) j
tk dtxij dxi ∧ dxi1 ∧ · · ·
i=1 j=1 0 ∂xi
/
.
∧ dx ∧ · · · ∧ dx
ij ik
.
The second term of the right hand side in this equality coincides with the second term
in Equation (1.3) except for the sign. Adding them together, we get
- 1
dH(ω) + Hdω = k tk−1 a(tx)dt dxi1 ∧ · · · ∧ dxik
i1 <···<ik 0
n -
1
∂a(tx)
+ i
tk
dtxi dxi1 ∧ · · · ∧ dxik
i1 <···<ik i=1
∂x 0
* +
- 1 k−1
n
k ∂a(tx) i
= kt a(tx) + t x dt dxi1 ∧ · · · ∧ dxik .
i1 <···<ik 0 i=1
∂xi
Notice that
n
∂a(tx) dk
ktk−1 a(tx) + tk xi i
= t a(tx) .
i=1
∂x dt
Thus, * +
- 1
n
∂a(tx) i
ktk−1 a(tx) + tk i
x dt
0 i=1
∂x
- 1
d k
= t a(t, x) dt = a(x).
0 dt
Then, we have
dH(ω) + Hdω = ai1 ···ik (x)dxi1 ∧ · · · ∧ dxik = ω,
i1 <···<ik
i.e.,
dH + Hd = id.
Therefore, the theorem is completed.
1 2
ωA (ξ) = (A, ξ), ωA (ξ, η) = (A, [ξ, η]), ∀ ξ, η ∈ R3 ,
where ( , ) stands for usual inner product and ( , [ ]) for triple scalar product.
1
Let A = A1 e1 + A2 e2 + A3 e3 and ωA = a1 dx1 + a2 dx2 + a3 dx3 . Then by
3
definition, on one hand, ωA (ej ) = ai dx1 (ej ) = Aj ; on the other hand, ωA (ej ) =
i=1
1
(A, ej ) = Aj . Thus, aj = Aj , i.e., ωA = A1 dx1 + A2 dx2 + A3 dx3 . Similarly, we
can get
2
ωA = A1 dx2 ∧ dx3 + A2 dx3 ∧ dx1 + A3 dx1 ∧ dx2 .
2
ωA =∗ (ωA
1
), 1
ωA = b (A).
Here, the top left hand corner “*” (“b”) represents the Hodge (sharp) operator respec-
tively, namely ∗ : ∧k (Rm ) → ∧n−k (Rm ); b : R → R∗ .
We now introduce three operators that play an important role in classical vector
analysis, i.e., gradient, curl, and divergence.
Definition 4.16. Let f ∈ C ∞ (R3 ) and A ∈ X (R3 ). The grad f and curl A ∈ X (R3 )
and div (A) ∈ C ∞ (R3 ) defined as follows
1 2 1 2 3
ωgradf
= df, ωcurl A
= dωA , and dωA = div A = ωA ,
By this definition,
1 ∂f ∂f ∂f
ωgrad f
= df = 1
dx1 + 2 dx2 + 3 dx3 .
∂x ∂x ∂x
Thus,
∂f ∂f ∂f
grad f = e1 + 2 e2 + 3 e3
∂x1 ∂x ∂x
∂f ∂f ∂f ∂f
= , , = ,
∂x1 ∂x2 ∂x3 ∂x
and so
88 1. Preliminaries of Differentiable Manifolds
2 1
ωcurl A
= dωA = d A1 (x)dx1 + A2 (x)dx2 + A3 (x)dx3
∂A1 1 ∂A ∂A
= 1
dx + 21 dx2 + 31 dx3 ∧ dx1
∂x ∂x ∂x
∂A2 1 ∂A ∂A
+ dx + 22 dx2 + 32 dx3 ∧ dx2
∂x1 ∂x ∂x
∂A3 1 ∂A ∂A
+ dx + 23 dx2 + 33 dx3 ∧ dx3
∂x1 ∂x ∂x
∂A3 ∂A ∂A1 ∂A
= − 32 dx2 ∧ dx3 + − 13 dx3 ∧ dx1
∂x2 ∂x ∂x3 ∂x
∂A2 ∂A
+ − 21 dx1 ∧ dx2 ,
∂x1 ∂x
where
∂A3 ∂A ∂A1 ∂A ∂A2 ∂A
curl A = − 32 e1 + − 13 e2 + − 21 e3
∂x2 ∂x ∂x3 ∂x ∂x1 ∂x
e1 e2 e3
∂
= ,
∂ ∂
1 ∂x2 ∂x3
∂x
A1 A2 A3
3 2
ωdiv A
= dωA = d(A1 dx2 ∧ dx3 + A2 dx3 ∧ dx1 + A3 dx1 ∧ dx2 )
∂A1 ∂A ∂A
= 1
+ 22 + 33 dx1 ∧ dx2 ∧ dx3 .
∂x ∂x ∂x
Therefore,
∂A1 ∂A ∂A
div A = + 22 + 33 .
∂x1 ∂x ∂x
Since
2 1
ωcurl grad f = d ωgrad f = d (d f ) = 0,
2
div curl (A)ω 3 = d ωcurl A
= d (d ωA1
) = 0,
we easily get two equalities in classical vector analysis:
∗ : Λp −→ Λn−p .
uΛ ∗ υ = (u, υ)en .
For brevity, we write uΛ ∗ υ as u ∗ υ.
If υ is a scalar, u must also be a scalar. By the above formula, we get ∗υ = υen .
Example 4.18. If
α= ai1 ···ip dxi1 ∧ · · · ∧ dxip ,
i1 <···<ip
then
∗α = bj1 ···jn−p dxj1 ∧ · · · ∧ dxjn−p ,
i1 <···<ip
where
bj1 ···jn−p = εi1 ···ip j1 ···jn−p ai1 ···ip ,
i1 <···<ip
Δ = d δ + δ d, Λk (M ) −→ Λk (M )
Δf = δdf.
∂f i
Δf = δdf = δ dx .
∂xi
The corresponding relationship in action on the form operators d and δ and action on
the coefficient of vector analysis can be summarized as follows:
1.5 Integration on a Manifold 91
d d d
form: Λ0
Λ1 Λ2 Λ3
δ δ δ
grad curl div
coef.: scalar vector vector scalar
−div curl −grad
We can easily obtain two equations in classical vector analysis:
d d = 0 : rot grad = 0,
div rot = 0.
δδ = 0 : −rot grad = 0,
−div rot = 0.
c = α1 c1 + · · · + αr cr , αi ∈ R, i = 1, · · · , r.
r
r
r
c1 + c2 = αi1 ci + αi2 ci = (αi1 + αi2 )ci ,
i=1 i=1 i=1
r
r
αc1 = α αi1 ci = (ααi1 )ci ,
i=1 i=1
92 1. Preliminaries of Differentiable Manifolds
r
where cj = αij ci (j = 1, 2) are two k-chains in M . Without loss of generality,
i=1
we assume that different chains c1 and c2 are generated by the same set of k-cubes
{c1 , · · · , cr }. For example, let c1 = c1 + 2c2 , c2 = c1 + c3 , where c3 = c2 . We only
need to rewrite c1 and c2 as c1 = c1 + 2c2 + 0 · c3 , c2 = c1 + 0 · c2 + c3 .
Boundary of Chains Corresponding to the exterior operator d: Ωk → Ωk+1 , there
is a boundary operator ∂ : C k (M ) → C k−1 (M ), as defined below:
k k
We call I(i,0) and I(i,1) as (i, 0)- and (i, 1)-surface respectively, and
k
∂I k = ∂[0, 1]k = (−1)i+α I(i,α)
n
.
i=1 α=0,1
c(i,α) = c ◦ I(i,α)
k
.
k
k
∂c = (−1)i+α c ◦ I(i,α)
k
= (−1)i+α c(i,α) .
i=1 α=0,1 i=1 α=0,1
The boundary of any k-chain c = αj cj is
j
∂c = αj ∂cj .
j
k
(I(i,α) k
)(j,β) (x) = I(i,α) (x1 , · · · , xj−1 , β, xj , · · · , xk−2 )
Similarly, we have
k
(I(j+1,β) k
)(i,α) (x) = I(j+1,β) (x1 , · · · , xi−1 , α, xi , · · · , xk−2 )
Thus, if i ≤ j, (I(i,α)
k k
)(j,β) (x) = (I(j+1,β) )(i,α) (x), it is easy to see that for any
k-cube c, (c(i,α) )(j,β) = (c(j+1,β) )(i,α) as i ≤ j. Now,
* k +
i+α
∂(∂c) = ∂ (−1) c(i,α)
i=1 α=0,1
k k−1
= (−1)i+α+j+β (c(i,α) )(j,β) .
i=1 α=0,1 j=1 β=0,1
In this sum, (c(i,α) )(j,β) and (c(j+1,β) )(i,β) occur simultaneously with the opposite
sign. Then, all terms disappear in pairs and ∂(∂c) = 0. Consequently, for any k-chain
r
c= αi ci , where ci (i = 1, · · · , r) are the k-cubes,
i=1
* +
r
r
∂(∂c) = ∂ ai ∂ci = ai ∂(∂ci ) = 0.
i=1 i=1
i.e., - -
ω= f dx1 · · · dxk ,
[0,1]k [0,1]k
In particular,
- -
f dx1 ∧ · · · ∧ dxk = (I k )∗ (f dx1 ∧ · · · ∧ dxk )
Ik [0,1]k
-
= f (x1 , · · · , xk ) dx1 · · · dxk .
[0,1]k
If c is a 0-cube, we define -
ω = ω(c(0)).
c
The integral of ω on a k-chain c = ai ci is
i
- -
ω= ai ω,
c i ci
4i ∧ · · · ∧ dxk .
f dx1 ∧ · · · ∧ dx
1.5 Integration on a Manifold 95
Notice that
-
k∗ 4i ∧ · · · ∧ dxk )
I(j,α) (f dx1 ∧ · · · ∧ dx
[0,1]k−1
- ∗ ∗
= f (x1 , · · · , xj−1 , α, xj , · · · , xk ) I(j,α)
k
dx1 ∧ · · · ∧ dxi ∧ · · · ∧ I(j,α)
k
dxk
[0,1]k−1
⎧
⎪ 0, i = j,
⎨
= -
⎪
⎩ 4i · · · dxk ,
f (x1 , · · · , α, · · · , xk ) dx1 · · · dx i = j.
[0,1]k−1
Thus,
-
4i ∧ · · · ∧ dxk
f dx1 ∧ · · · ∧ dx
∂I k
n -
= (−1)j+α k∗
I(j,α) 4i ∧ · · · ∧ dxk )
(f dx1 ∧ · · · ∧ dx
j=1 α=0,1 [0,1]k−1
-
= (−1)i+1 f (x1 , · · · , xi−1 , 1, · · · , xk )
[0,1]k−1
−f (x1 , · · · , 0, · · · , xk ) d x1 ∧ · · · ∧ d.
xi ∧ · · · ∧ d xk .
In other words, - -
dω = ω.
Ik ∂I k
k
∂c = (−1)i+α c(i,α)
i=1 α=0,1
k
= (−1)i+α c ◦ I(i,α)
k
,
i=1 α=0,1
by definition of integration,
-
k -
i+α
ω= (−1) ω
∂c k
c◦I(i,α)
i=1 α=0,1
k - -
∗
= (−1) i+α
c ω= c∗ ω,
∗
I(i,α) ∂I k
i=1 α=0,1
- - - -
dω = c∗ dω = dc∗ ω = c∗ ω,
c Ik Ik ∂I k
- -
and so dω = ω, for any singular k-cube c.
c ∂c
k
Finally, if c is a k-chain, i.e., c = αi ci , where ci are singular k-cubes, then
i=1
-
k -
k - -
dω = ai dω = ai ω= ω.
c i=1 ci i=1 ∂ci ∂c
ω 1 = p1 dq1 + · · · + pn dqn ,
Proof. Since
∂Q ∂P
d (P dx + Qdy) = − dx ∧ dy,
∂x ∂y
- -
the Stokes theorem ωdl = dω implies
∂c c
- -
∂Q ∂P
− dxdy = P dx + Qdy,
D ∂x ∂y l
- -
2 ∂Ax ∂Ay ∂Az
dωA = + + dx ∧ dy ∧ dz
c3 c3 ∂x ∂y ∂z
-
= div Adxdydz,
D
- -
ω = Ax dydz + Ay dxdz + Az dxdy.
∂c3 S
Zk ker dk
H k (M, R) = k
≡ ,
B imdk−1
Bk ⊆ Z k .
1.7 Lie Derivative 99
Similarly, the set of all k-cycles, denoted by Zk , and the set of all k-boundaries,
denoted by Bk , form a subspace of the vector space C k . The corresponding quotient
space
Zk ker ∂k
Hk (M, R) = =
Bk im∂k+1
is called the k-th homology space. An element in Hk is a class of cycles differing from
one another only by a boundary.
Definition 6.1. The dimension of H k (Hk ), denoted by bk (bk ), is called the k-th Betti
number.
Theorem 6.2 (De Rham theorem). The two Betti numbers are the same, i.e.,
bk = bk .
Z 0 (M ) = R × · · · × R .
0 12 3
m
i=0
100 1. Preliminaries of Differentiable Manifolds
Rn , where , is the dual bracket between Tx Rn and Tx∗ Rn . This defines the natural
bilinear mapping of
Λ1 (Rn ) × X (Rn ) −→ C ∞ (Rn ).
∂f
If ω 1 = df = dxi , then
∂xi
n
∂f n
∂f
df, X = X
i i
= Xi (x) i .
i=1
∂x i=1
∂x
∂f
n
Denote df, X = LX f , i.e., LX f = i
, ∀ f ∈ C ∞ (Rn ). Thus, any smooth
Xi
i=1
∂x
vector field may be viewed as a linear partial differential operator on Rn of order 1,
without zero terms and smooth coefficient, i.e., there is a correspondence between
X(x) ∈ X (Rn ) and LX :
n
n
∂
X(x) = Xi (x)ei −→ LX = Xi (x) .
i=1 i=1
∂xi
n
It is one to one, and so hereafter, we can also write X(x) = Xi (x)ei as
i=1
n
∂
X(x) = Xi (x) .
i=1
∂xi
[X, Y ] = XY − Y X,
i.e.,
[X, Y ]f = X(Y f ) − Y (Xf ), ∀ f ∈ C ∞ (R),
n
∂ n
∂
Proposition 7.2. Let X = Xi (x) i
,Y = Yi (x) i , then
i=1
∂x i=1
∂x
n
∂Xi ∂
n
∂Yi
[X, Y ] = Xk − Yk .
i=1 k=1
∂xk ∂xk ∂xi
Proof. ∀ f ∈ C ∞ (Rn ),
1.7 Lie Derivative 101
Thus,
n
∂Xi ∂
n
∂Yi
[X, Y ] = Xk − Yk .
i=1 k=1
∂xk ∂xk ∂xi
The i-th component of [X, Y ] is
n
∂Yi ∂Xi
[X, Y ]i = Xk − Yk .
k=1
∂xk ∂xk
(1) Phase space: The set of all possible states of a process is called phase space. For
example, consider the motion of system in classical mechanics, whose future and past
are uniquely determined by the initial position and initial velocities of all points in the
system. The phase space of a mechanical system is the set whose elements are the sets
of positions and velocities of all points of the given system. This is exactly the tangent
space we discussed earlier.
(2) Differentiable process: A process is called differentiable if its phase space has
the structure of a differentiable manifold, and the change of state with time is described
by differentiable functions. Let M be the phase space. A point of this space is a defined
state of the process. Assume that at instant t = 0 the process was in state x. Then at
another moment, the process t will be in another state. Denote this new state of the
process by g t x. We have defined for each t a mapping g t
g t : M −→ M.
This is called the transformation in time t, which takes the state at instance 0 to the
state at the instant t
g 0 = id, g t+s = g t · g s .
Let y = g s x be the state after time s, and the state z = g t y again after time t. The
effect is the same as advancing x to time t + s, i.e.,
z = g t+s x.
g t+s = g t · g s , g 0 = I.
Definition 7.5. The set M and the corresponding one parameter transformation group
{g t } that maps M to itself compose (M, {g t }), which is called the phase flow, where
set M is called the phase space, and its element is called the phase point.
d x(t)
= X(x(t)), x|t=0 = x0 (initial value).
dt
The image of the mapping x(t) is called a phase space, and the graph of the map-
ping x(t) is called an integral curve. The integral curve lies in the direct product of
the time axis and the phase space, which is called M × R extended space. Such an
equation is called a dynamic system, whose phase flow g t , M → M, {g t , t ∈ R}
composes a group:
Definition 7.6. Let X and Y be two vector fields. The Lie derivative LX Y of Y with
respect to X is defined by
d t∗
LX Y = φ Y t=0 ,
dt X
where φtX is the flow of X.
104 1. Preliminaries of Differentiable Manifolds
= (Dφ−t
X (y)) df (x), Y (y)
= d(f ◦ φ−1
X (y)) , Y (y)
= df"(y), Y (y) f" = f ◦ φtX
d t∗ d n
∂f
φX f = f φtX (x) = Xi φtX (x)
dt dt ∂xi
i=1
∗
= (Xf ) φtX (x) = φtX (Xf ).
3◦ The proof is as follows:
d t∗ d t∗ ∗
(LX Y )f = φX Y f t=0 = φX (Y · φ−t
X f ) t=0
dt dt
∗ ∗
=
∗
φtX XY · φ−t
∗
− φtX Y φ−t
X f X Xf t=0
= XY f − Y Xf = [X, Y ]f, ∀ f.
Thus, we get
LX Y = [X, Y ].
Therefore, the theorem is completed.
By the equality 3◦ , the Jacobi identity about the Poisson bracket { , } shows that
the operator LX is a { , }-derivative on the algebra X (Rn ) with binary operator [ , ],
i.e.,
LX {X1 , X2 } = {LX X1 , X2 } + {X1 , LX X2 }.
Definition 7.8. ∀ω ∈ Ωk (Rn ), the Lie derivative LX ω of ω with respect to a vector
field X ∈ X (Rn ) is defined by
d ∗ 1 ∗
LX ω = φtX ω = lim φtX ω − ω .
dt t=0 t→0 t
1.7 Lie Derivative 105
Theorem 7.9. The Lie derivative LX with respect to X ∈ X (Rn ) has the following
properties:
n
∂f
1◦ LX f = Xf = Xi i , f ∈ C ∞ (Rn ), i.e., the Lie derivatives of a
i=1
∂x
function f with respect to X is the directional derivative in direction X(x).
2◦ LX is a Λ-derivative, i.e., LX is R-linear,
3◦ LX d = d LX .
d t∗ ∗
Proof. 1◦ We have φ f = φtX (Xf ), and so
dt X
d
(φt f ) = (φtX Xf )t=0 = Xf.
∗ ∗
LX f =
d t t=0 X
= LX ω1 ∧ ω2 + ω1 ∧ LX ω2 .
In the first part, it must be σ(1) = 1 since σ(1) < · · · < σ(k). Similarly, in the second
part, σ(k + 1) = 1.
Set
Then,
Thus,
1.7 Lie Derivative 107
· ω3 (ξσ(k+1) , · · · , ξσ(k+l) )
+ (−1)k ε(σ )ω1 (ξσ(1) , · · · , ξσ(k) )
σ(1)<···<σ(k), σ(k+1)<···<σ(k+l)
= ε(σ )(iX ω1 )(ξσ (1) , · · · , ξσ (k−1) )
σ (1)<···<σ (k), σ (k−1)<···<σ (k+l−1)
+ (−1)k ε(σ )ω1 (ξσ (1) , · · · , ξσ (k) )(iX ω3 )
σ (1)<···<σ (k), σ (k+1)<···<σ (k+l−1)
iX (ω1 ∧ ω3 ) = iX ω1 ∧ ω3 + (−1)k ω1 ∧ iX ω3 .
= iX dω ∧ df + (−1)k+1 dω ∧ iX df + diX ω ∧ df
LX = iX d + diX .
Lf X = (dif X + if X d)ω
= d(f iX ω) + f iX dω
= df ∧ iX ω + f diX ω + f iX dω
= (f LX + df ∧ iX )ω,
so
Lf X = f LX + df ∧ iX .
Therefore, the theorem is completed.
It is linear, skew-symmetric.
Proof.
1◦ By the definition of LX and the Theorem 7.7, since
1.7 Lie Derivative 109
∗
(φtX ω)(X1 , · · · , Xk ) = ω(φtX ω)(φtX∗ X1 , · · · , φtX∗ Xk )
∗ ∗
= ω(φ−t −t
X X1 , · · · , φX Xk )(φX X)
t
∗ ∗ ∗
= φtX (ω(φ−t −t
X X1 , · · · , φX Xk )),
we have
d ∗
(LX ω)(X1 , · · · , Xk ) = (φt ω)(X1 , · · · , Xk )
d t t=0 X
t∗
d ∗
−t∗
= φ ω(φ−t
X X1 , · · · , φX Xk )
d t t=0 X
k
= LX ω(X1 , · · · , Xk ) − ω(X1 , · · · , LX Xi , · · · , Xk ).
i=1
k
= LX0 ω(X1 , · · · , Xk ) − ω(X1 , · · · , LX0 Xi , · · · , Xk )
i=1
−(diX0 ω)(X1 , · · · , Xk ),
(diX0 ω)(X1 , · · · , Xk )
k
= i , · · · , Xk )
(−1)i−1 LXi iX0 ω(X1 , · · · , X
i=1
+ i , · · · , X
(−1)i+j iX0 ω(LXi Xj , X1 , · · · , X j , · · · , Xk )
1≤i≤j≤k
k
= i , · · · , Xk )
(−1)i−1 LXi ω(X0 , X1 , · · · , X
i=1
+ i , · · · , X
(−1)i+j−1 ω(LXi Xj , X0 , · · · , X j , · · · , Xk ).
1≤i≤j≤k
Thus, we get
110 1. Preliminaries of Differentiable Manifolds
d ω(X0 , · · · , Xk )
k
= LX0 ω(X1 , · · · , Xk ) + (−1)j ω(X1 , · · · , LX0 Xj , · · · , Xk )
j=1
k
+ i , · · · , Xk )
(−1)i LXi ω(X0 , · · · , X
i=1
+ i , · · · , X
(−1)i+j ω(LXi Xj , X0 , · · · , X j , · · · , Xk )
1≤i<j<k
k
= i , · · · , Xk )
(−1)i LXi ω(X0 , · · · , X
i=0
+ i , · · · , X
(−1)i+j ω(LXi Xj , X0 , · · · , X j , · · · , Xk ).
i<j
[AA88] D.V. Anosov and V.I. Arnold: Dynamical Systems I. Springer, Berlin, (1988).
[AA89] V. I. Arnold and A. Avez: Ergodic Problems of Classical Mechanics. Addison-Wesley
and Benjamin Cummings, New York, (1989).
[Abd99] S. S. Abdullaev: A new integration method of Hamiltonian systems by symplectic
maps. J. Phys. A: Math. Gen., 32(15):2745–2766, (1999).
[Abd02] S. S. Abdullaev: The Hamilton–Jacobi method and Hamiltonian maps. J. Phys. A:
Math. Gen., 35(12):2811–2832, (2002).
[AKN78] V. I. Arnold, V. V. Kozlov, and A. I. Neishtadt: Mathematical Aspects of Classical
and Celestial Mechanics. Springer, Berlin, Second edition, (1978).
[AM78] R. Abraham and J. E. Marsden: Foundations of Mechanics. Addison-Wesley, Reading,
MA,Second edition, (1978).
[AMR88] R. Abraham, J. E. Marsden, and T. Ratiu: Manifolds, Tensor Analysis, and Applica-
tions. AMS 75. Springer-Verlag, Berlin, Second edition, (1988).
[AN90] A. I. Arnold and S.P. Novikov: Dynamical System IV. Springer Verlag, Heidelberg,
(1990).
[AP92] D. K. Arrowsmith and C. M. Place: Dynamical Systems: Differential Equations, Maps,
and Chaotic Behavior. Chapman & Hall, New York, (1992).
[Arn78] V. I. Arnold: Ordinary Differential Equations. The MIT Press, New York (1978).
[Arn88] V. I. Arnold: Geometrical Methods in the Theory of Ordinary Differential Equations.
Springer-Verlag, Berlin, (1988).
[Arn89] V. I. Arnold: Mathematical Methods of Classical Mechanics. Berlin Heidelberg:
Springer-Verlag, GTM 60, Berlin ,Second edition, (1989).
[Ber00] R. Berndt: An Introduction to Symplectic Geometry. AMS Providence, Rhode Island,
(2000).
[Bir23] G. D. Birkhoff: Relativity and Modern Physics. Harvard Univ. Press, Cambridge,
Mass., Second edition, (1923).
[BK89] G.W. Bluman and S. Kumei: Symmetries and Differential Equations. AMS 81.
Springer-Verlag, New York, (1989).
[Car65] C. Carathe’odory: Calculus of Variation and Partial Differential Equations of First
Order, Vol.1. Holden-Day, San Franscisco, (1965).
[Car70] H. Cartan: Differential Forms. Houghton-Mifflin, Boston, (1970).
[Che53] S. S. Chern: Differential Manifolds. Lecture notes. University of Chicago, (1953).
[Ede85] D. G. B Edelen: Applied Exterior Calculus. John Wiley and Sons, New York, First
edition, (1985).
[Fla] H. Flanders: Differential Forms. Academie Press, New York, Second edition. (1963).
[GS84] V. Guillemin and S. Sternberg: Symplectic Techniques in Physics. Cambridge Univer-
sity Press, Cambridge, (1984).
[Lan95] S. Lang: Differential and Riemannian Manifolds. Springer-Verlag, New York, (1995).
[LM87] P. Libermann and C.M. Marle: Symplectic Geometry and Analytical Mechanics. Rei-
del Pub. Company, Boston, First edition, (1987).
[Mac70] S. MacLanc: Hamiltonian mechanics and geometry. Amer. Math. Mon., 77(6):570–
586, (1970).
112 Bibliography
[Poi93] H. Poincaré: Les Méthodes Nouvelles de la Mécanique Céleste, Tome II. Gauthier-
Villars, Paris, Second edition, (1893).
[Poi99] H. Poincaré: Les Méthodes Nouvelles de la Mécanique Céleste. Tome III. Gauthiers-
Villars, Paris, Second edition, (1899).
[Sc77] M. Schreiber: Differential Forms. Springer-Verlag, New York, First edition, (1977).
[Sie43] C.L. Siegel: Symplectic geometry. Amer.and math. J. Math, 65:1–86, (1943).
[Spi68] M. Spivak: Calculus on Manifolds. The Benjamin/Cummings publishing company,
London, New York, First edition, (1968).
[Tre75] F. Treves: Pseodo-Differential Operator. Acad. Press, New York, First edition, (1975).
[Wei77] A. Weinstein: Lectures on symplectic manifolds. In CBMS Regional Conference, 29.
American Mathematical Society, Providence, RI, (1977).
[Wes81] C. Von. Westenholz: Differential Forms in Mathmatical Physics. North-Holland,
Amsterdam, Second edition, (1981).
Chapter 2.
Symplectic Algebra and Geometry
Preliminaries
n
ϕA (x, y) = x Ay = aij xi yj .
i,j=1
i.e.,
Proof. The sufficiency is trivial. We only need to prove the necessity. Without loss of
generality, we can assume F = R. Then, we have
ker(A) = {y ∈ Rn | Ay = 0} = {y ∈ Rn | x Ay = 0, ∀ x ∈ Rn },
ker(B) = {y ∈ Rn | By = 0} = {y ∈ Rn | x By = 0, ∀ x ∈ Rn }.
Av = 0 ⇐⇒ Bv = 0.
{Av} = {Bv}.
Then,
μ1 Bv1 + · · · + μr Bvr = μB(v1 + v2 + · · · + vr )
= μBv1 + · · · + μBvr .
After manipulation, we get (μ1 − μ)Bv1 + · · · + (μr − μ)Bvr = 0, i.e.,
μ = μ1 = · · · = μr .
Therefore,
Avi = μBvi , i = 1, · · · , r.
Similarly for i = r + 1, · · · , n, Avi = 0 = μBvi . Thus, we have obtained A = μB.
C z = x + i y → z = x − i y ∈ C,
such that
z1 · z2 = z1 · z2 , z1 + z2 = z1 + z2 .
This leads to a new kind of binary forms.
Definition 1.15 (Sesquilinear). A sesquilinear form on Cn is a mapping φ : Cn ×
Cn → C, such that for all u, v, x ∈ Cn , α, β ∈ C, we have
1◦ φ(αu + βv, x) = α φ(u, x) + β φ(v, x).
2◦ φ(x, αu + βv) = α φ(x, u) + β φ(x, v).
Similarly, there exists a 1-1 correspondence between the complex matrix space M (n, C)
and the space of sesquilinear forms on Cn .
In fact, a complex matrix A ∈ M (n, C) has a natural correspondence to a
sesquilinear form φA , which satisfies
n
φA (x, y) = x∗ Ay = aij xi yj .
i,j
or
x∗ Ay = 0 ⇐⇒ x∗ A∗ y = 0.
n
(x, y) = (x, y)I = x y = xi yi , I = I.
i=1
n
This induces a Euclidean length measure |x|2 = (x, x) = x2i .
i=1
n
[x, y] = [x, y]J = x Jy = (xi yn+i − xn+i yi ),
i=1
where !
0 In
J= = J2n , J = J −1 = −J.
−In 0
As n = 1, we have
x y1
[x, y] = x1 y2 − x2 y1 = 1 ,
x2 y2
which represents the oriented area of the parallelogram formed by the vector x, y in
R2 (see the image below).
(x2 , y2 )
6
y
For general n, we get
n : x
xi yi
[x, y] = x Jy = -
xn+i yn+i O (x1 , y1 )
i=1
Remark 1.24 (Pfaffian theorem). For any n, there exists a polynomial Pn (xij ) with
integer coefficients in variables xij (i < j) such that
2
det K = det [kij ] = [Pn (kij )] , ∀ anti-symmetric matrix K.
A ∈ O(n, F) ⇐⇒ A IA = A A = I.
A ∈ Sp(n, F) ⇐⇒ Ā JA = J.
A ∈ U (n, C) ⇐⇒ A∗ IA = I.
O(S, n, F) = G(S, n, F)
( )
= A ∈ GL(n, F) | (Ax, Ay)S = (x, y)S , ∀ x, y ∈ Fn
(
= A ∈ GL(n, F) | A SA = S}
( )
= A ∈ GL(n, F) | A−1 = S −1 A S .
¯ n, F) ≡ O(n, F).
Special case: O(I,
Anti-symmetrical case: G = K, K = −K, ϕK (x, y) = [x, y]K = x Ky, where
G is called a K-symplectic group.
Property 1.26.
1◦ ρ(z + w) = ρ(z) + ρ(w), ∀z, w ∈ Cn .
2◦ ρ(αz) = αρ(z), ∀α ∈ R.
! ! !
−y 0 −I x
3◦ ρ(iz) = ρ(−y + ix) = = = −Jρ(z).
x I 0 y
4◦ ρ((α + iβ)z) = (αI − βJ)ρ(z), α + iβ ∈ C.
5◦ ρ(0) = 0 ∈ R2n , 0 ∈ Cn .
!
A −B
For C = A + iB ∈ M (n, C), set R(C) = ∈ M (2n, R).Similarly,
B A
R : C → R(C), M (n, C) → M (2n, R) is injective.
Assume C = A + iB ∈ M (n, C), w = Cz. Then,
i.e., ! ! !
u A −B x
= ,
v B A y
or
ρ(w) = R(C)ρ(z) = ρ(Cz).
Analogously, R satisfies the following properties:
Property 1.27.
1◦ R(On ) = O2n , On ∈ M (n, C).
2◦ R(In ) = I2n , In ∈ M (n, C).
3◦ R(αC) = αR(C), ∀ α ∈ R.
!
−B −A
4◦ R( i C) = R( iA − B) = = −JR(C).
A −B
5◦ R(C1 + C2 ) = R(C1 ) + R(C2 ), ∀ C1 , C2 ∈ M (n, C).
6◦ R(C1 · C2 ) = R(C1 )R(C2 ).
Theorem 1.28.
A −B
det(A + iB) = 0 ⇐⇒ 0.
=
B A
122 2. Symplectic Algebra and Geometry Preliminaries
The above equation shows that the Hermitian scalar product of z and w, w, zH
consists of two parts: its real part is a Euclidean scalar product in R2n (whose measure
is a symmetric matrix S), denoted by the round bracket and its imaginary part can be
taken as a new scalar product in R2n (whose measure is an anti-symmetric matrix K),
denoted by the square bracket. Therefore, we have w, zH =(W, Z)S + i [W, Z]K ,
where !
u
W = ρ(w) = ,
v
!
x
Z = ρ(z) = ∈ R2n ,
y
H = P + i Q, P = P,
!
P −Q
Q = −Q, S = S = ,
Q P
!
Q P
K = −K = .
−P Q
l : V −→ V, u −→ i u = l(u).
T (i u) = i T (u) ∼ T l = lT.
Let
We can easily see that V is complexifiable iff dimV (R) = 2n. This is because the
operator equation l2 = −I in L(V, R) corresponds to the matrix equation X 2 = −I
in M (m, R), m =dimR (V ), which has no real solution for m = 2n + 1. (Since
from X 2 = −I it follows that all eigenvalues of X are ±i, while for m = 2n + 1,
X ∈ M (m, R) has at least one real eigenvalue.) When m = 2n, there is a special
solution
X = ±J, (±J)2 = −I.
If we introduce an isomorphism J on R2n , which satisfies J 2 = −I, then we
say that R2n is equipped with the complex structure. Hence, we can define operation
(a + ib)u = au + bJu, and R2n becomes the complex n-dimensional space. Cn is
called R2n complexifiable.
Definition 1.32. A Lie algebra is a vector space equipped with a binary operation
L × L → L, which satisfies the Jacobi identity.
Hence, M (n, F), equipped with the above commutator, becomes a Lie algebra,
denoted as gl(n, F). Since gl(n, F) is the tangent vector space to GL(n, F) at I,
gl(n, F) is called Lie algebra of the Lie group GL(n, F).
Definition 1.33. The Lie algebra of the Lie group SL(n, F) is defined as follows:
sl(n, F) = {B ∈ gl(n, F) | trB = 0}.
Remark 1.34. If trB1 = trB2 = 0, then tr{B1 , B2 } = tr(B1 B2 − B2 B1 ) = 0.
Therefore, sl(n, F) is closed under { , }. As the matter of fact, for any A, B, tr{A, B}
is always equal to 0.
Definition 1.35. The Lie algebra of the Lie group G(G, n, F) is defined as follows:
g(G, n, F) = {B ∈ gl(n, F)|B G + GB = 0}.
Remark 1.36. g(G, n, F) is closed under { , }, i.e.,
{B1 , B2 } ∈ g(G, n, F), ∀ B1 , B2 ∈ g(G, n, F).
If Bi G = −GBi (i = 1, 2), then
{B1 , B2 } G = (B1 B2 − B2 B1 ) G
= (B2 B1 − B1 B2 )G
= B2 B1 G − B1 B2 G
= B2 (−GB1 ) − B1 (−GB2 )
= G(B2 B1 − B1 B2 )
= −G{B1 , B2 }.
∞
Bk
Proof. The series is uniformly convergent in ||B|| ≤ ρ (ρ > 0). Therefore,
k!
k=0
Fij = (exp B)ij are the entire analytic functions of the complex variables bkl (k, l =
1, · · · , n),
Fij (bkl ) = δij + bij + · · · + (terms of degree) ≥ 2.
Therefore, the Jacobian matrix is
∂(expB)ij ∂Fij (bkl ) kl
= = δij ,
∂ b̄kl ∂bkl
'
kl 1, k = i, l = j,
where δij =
0, otherwise.
kl
Since det (δij ) = 1 = 0, by implicit function theorem, there exists a neighborhood
W ⊂ {||B|| < ρ}, such that B → exp B is a diffeomorphism on W .
Proposition 1.41. The function A(t) = exp (tB), ∀ t ∈ R satisfies the following prop-
erties:
A(0) = I,
A(t1 + t2 ) = A(t1 )A(t2 ) = A(t2 )A(t1 ),
A(−t) = (A(t))−1 ,
d d
A(t) = BA(t), A(t)|t=0 = B.
dt dt
Proposition 1.42. For A1 , A2 ∈ GL(n, C), its commutator is defined by
{A1 , A2 }G = A1 A2 A−1 −1
1 A2 (commutator in Lie group).
Then for Ai (t) = exp tBi (i = 1, 2), ∀ t ∈ R, we have
{A1 (t), A2 (t)}G = I + t2 {B1 , B2 }g + o(t4 ),
{A1 (t), A2 (t)}G |t=0 = In ,
d
{A1 (t), A2 (t)}G |t=0 = 0,
dt
1 d2 ( )
A1 (t), A2 (t) G |t=0 = {B1 , B2 }g ,
2 d t2
where {B1 , B2 }g = B1 B2 − B2 B1 is the commutator in Lie algebra.
n
Proposition 1.43. If B ∈ g(G, n, F) and f (λ) = αk λk , αk ∈ F, then (B k ) G =
k=0
G(−B)k , (f (B)) G = Gf (−B).
Theorem 1.44. A(t) = exp (tB) ∈ G(G, n, F), ∀ t ∈ R iff B ∈ g(G, n, F).
Proof. Let C(t) = A (t)GA(t), then C(0) = G,
d d A (t) d A(t)
C(t) = GA(t) + A (t)G
dt dt dt
= BA (t)GA(t) + A (t)GBA(t) (1.1)
= A (t)(B G + GB)A(t).
d
Thus, A(t) ∈ G(G, n, F), ∀t ∈ R iff C(t) ≡ G, i.e., C(t) = 0. Then, in order
dt
to prove the theorem, we need only to show that the latter condition is equivalent to
B ∈ g(G, n, F), i.e., B G + GB = 0.
d d
If C(t) = 0, then C(t)t=0 = I (B G + GB)I = B G + GB = 0, and so
dt dt
B ∈ g(G, n, F). Conversely, if B G + GB = 0, by (1.1),
d
C(t) = 0,
dt
then C(t) = C(0) = G, i.e., A (t)GA(t) = G, ∀ t ∈ R.
128 2. Symplectic Algebra and Geometry Preliminaries
{B1 , B2 } ∈ cg(G, n, F). This shows that cg(G, n, F) is closed under { , }. There-
fore, cg(G, n, F) is a subalgebra of gl(n, F) equipped with the induced binary opera-
tion { , }, called as the Lie algebra of the conformally invariant group CG(G, n, F).
Theorem 1.46. Let A(t)= exp(tB), then A(t) ∈ CG(G, n, F) iff B ∈ cg(G, n, F).
The representative matrix G will change as the base changes. In this section, we want
to make sure how the matrix G changes.
Let Fn = {e1 , · · · , en } = {f1 , · · · , fn }, where e1 , · · · , en is the standard base and
f1 , · · · , fn is a new base. Then,
n
fi = tji ej , T = [tij ] ∈ GL(n, F),
j=1
⎡ ⎤ ⎡ ⎤
x1 u1
n
n
⎢ ⎥ ⎢ .. ⎥ ,
x= xi ei = ui f¯i , x = ⎣ ... ⎦ = T u = T ⎣ . ⎦
i=1 i=1 xn un
⎡ ⎤ ⎡ ⎤
y1 v1
n
n
⎢ ⎥ ⎢ .. ⎥ .
y= yi ei = vi f¯i , y = ⎣ ... ⎦ = T v = T ⎣ . ⎦
i=1 i=1 yn vn
2.2 Canonical Reductions of Bilinear Forms 129
Assume that under the new base f¯1 , · · · , f¯n , the representative matrix of ϕ is G. Then,
%
n
n &
Gij = ϕ(f¯i , f¯j ) = ϕ tki ek , tlj el
k=1 l=1
n
= tki ϕ(ek , el )tlj = (T GT )ij ,
k,l=1
i.e.,
G = T GT,
n
n
n
n
ϕ(x, y) = ϕ xi ei , yj ej = xi ϕ(ei , ej )yj = xi Gij yj = x Gy
i=1 j=1 i,j=1 i,j=1
n
n
n
= ϕ ui f¯i , vj f¯j = ui ϕ(f¯i , f¯j )vj
i=1 j=1 i,j=1
n
= ui Gij vj = u Gv.
i,j=1
n
If f¯1 , · · · , f¯n is another base on Cn and f¯j = tij ei , T = [tij ] ∈ GL(n, C),
i=1
then similarly we can get
G = T ∗ GT,
where G is the representative matrix of φ under the base f¯1 , · · · , f¯n .
130 2. Symplectic Algebra and Geometry Preliminaries
Remark 2.3. G is a conformal Hermitian matrix, i.e., G = εG with G, ε ∈ C, and
iθ
|ε| = 1, if and only if ε = eiθ and G∗ = eiθ G, if and only if e 2 G is a Hermitian
matrix.
U ϕ = {x ∈ Rn /Cn | ϕ(x, y) = x Gy = 0, ∀y ∈ U }
Thus, the canonical forms listed in Table 2.1 have the following alternative forms:
2.2 Canonical Reductions of Bilinear Forms 131
G = G, in C, ∃ u ∈ Cr , s.t.
⎡ ⎤ {a1 } ∩ {a1 }ϕ = {0},
. ϕ(u, u) = 0.
⎢ Ir
.
. ⎥ {a1 }ϕ = {a2 , · · · , ar },
⎢ ⎥ u 5 6
⎢ ... ... ⎥ Let a1 = 1 0
⎢ ⎥ ϕ(u, u)
⎣ ⎦ G∼
. 0 G1
.
. On−r =⇒ ϕ(a1 , a1 ) = 1
∃ u ∈ Rr , s.t. {a1 } ∩ {a1 }ϕ = {0},
G = G, in R,
⎡ ⎤ ϕ(u, u) = 0.
Ip {a1 }ϕ = {a2 , · · · , ar },
⎢ ⎥ u
⎣ −Iq ⎦ Let a1 = ⎡ ⎤
|ϕ(u, u)| ±1 0
On−r G∼ ⎣ ⎦
p+q =r =⇒ ϕ(a1 , a1 ) = ±1 0 G1
G∗ = G, G∗ = eiθ G ∃ u ∈ Cr , s.t.
⎡ ⎤
Ip
⎢ ⎥ ϕ(u, u) = 0. {a1 } ∩ {a1 }ϕ = {0},
⎣ −Iq ⎦, u
Let a1 =
On−r
|ϕ(u, u)| {a1 }ϕ = {a2 , · · · , ar },
⎡ ⎤ ⎡ ⎤
−iθ =⇒ ϕ(a1 , a1 ) ±1 0
⎢ e 2 Ip ⎥ G∼ ⎣ ⎦
⎢ −iθ ⎥
⎢ −e 2 ⎥ = signϕ(u, u) 0 G1
⎣ Iq ⎦
On−r
= ±1
p+q =r
⎡ ⎤
0 Ir
⎣ −Ir 0 ⎦ ∼ G = −G, in R or in C,
On−r
!
Ir
∼ G = G, in C,
On−r
⎡ ⎤
0 Is
⎢ Is 0 ⎥
⎣ ⎦ ∼ G = G, in R or (G∗ = G, in C ),
σId
On−r
⎡ ⎤
0 Is
⎢ e−iθ I ⎥
⎢ 0 ⎥ ∗
⎥ G = eiθ G,
s
⎢ iθ in C,
⎣ σe− 2 Id ⎦
On−r
132 2. Symplectic Algebra and Geometry Preliminaries
Theorem 2.6 (Sylvester’s law of inertia). Let ϕ(x) be a quadratic form in Rn and
x = T y, det(T ) = 0. If
then p = q.
Similarly, let φ(x) be a quadratic form in Cn and x = T y, det T = 0. If
then p = q.
aa + c∗ G1 c = 1, ab̄ + c∗ G1 d = 0,
b̄a + d∗ G1 c = 0, b̄b̄ + d∗ G1 d = G2 .
Let dλ = d + λcb ,
We denote ⎡ ⎤
..
⎛ ⎞ .
⎢ ±1 ⎥
⎢ .. ⎥
⎢ ⎜ .. ⎟ . ⎥
⎢ ⎝ . ⎠ 0 ⎥
⎢
$1 = ⎢ .. ⎥
G ±1 . ⎥,
⎢ r−1 ⎥
⎢ ··· ··· ··· .. ··· ⎥
⎢ . ⎥
⎣ 0 G1 ⎦
..
.
134 2. Symplectic Algebra and Geometry Preliminaries
⎡ ⎤
..
⎛ ⎞ .
⎢ ±1 ⎥
⎢ .. ⎥
⎢ ⎜ .. ⎟ . ⎥
⎢ ⎝ . ⎠ 0 ⎥
⎢
$2 = ⎢ .. ⎥
G ±1 r−1 . ⎥,
⎢ ⎥
⎢ · · · · · · ··· .. ··· ⎥
⎢ . ⎥
⎣ 0 G2 ⎦
..
.
$1 ∼
then by the result just proved above, G
c $
G2 . By recursions, we can finally get
c
G1 ∼ G2 .
!
Ip−p 0
= −T ∗ T.
0 −In−p
1 = −(|ti1 |2 + · · · + |tn−p |2 ),
which is a contradiction.
For comparison, we list the canonical forms of Hermitian, conformal Hermitian ma-
trices, real symmetric, and anti-symmetric matrices under unitary or orthogonal trans-
formations in the Table 2.2. The content can be found in any standard textbook.
2.2 Canonical Reductions of Bilinear Forms 135
Next, we consider Jordan canonical forms. Let us first recall the Jordan canonical form
for a general real matrix A ∈ M (n, R). A Jordan canonical form viewed in real space
is different from the Jordan canonical form viewed in complex space.
1. Elementary divisors in complex space
In complex space, the elementary divisor corresponding to a paired-complex conjugate
eigenvalue α ± iβ, β = 0 is of the form
[λ − (α + iβ)]p , [λ − (α − iβ)]p .
The corresponding Jordan blocks are
⎡ ⎤ ⎡ ⎤
α + iβ 1 α − iβ 1
⎢ .. ⎥ ⎢ .. ⎥
⎢ α + iβ . ⎥ ⎢ α − iβ . ⎥
⎢ ⎥ , ⎢ ⎥ .
⎢ .. ⎥ ⎢ .. ⎥
⎣ . 1 ⎦ ⎣ . 1 ⎦
α + iβ p×p
α − iβ p×p
(λ − γ)q .
The elementary divisor and the Jordan block corresponding to a real eigenvalue γ
is the same as in complex space, i.e.,
2.3 Symplectic Space 137
⎡ ⎤
γ 1
⎢ .. ⎥
⎢ γ . ⎥
⎢ ⎥ ∼ (λ − γ)q .
⎢ .. ⎥
⎣ . 1 ⎦
γ q×q
U ⊂ V ⇐⇒ U ⊃ V , U ⊂ V ⇐⇒ U ⊥ ⊃ V ⊥ ;
(U ∩ V ) = U + V , (U ∩ V )⊥ = U ⊥ + V ⊥ ;
(U + V ) = U ∩ V , (U + V )⊥ = U ⊥ ∩ V ⊥ ;
∃ U, U ∩ U = {0}, U ∩ U ⊥ = {0};
U + U = R2n , U + U ⊥ = R2n .
⇐⇒ V ∩ V = V 0 = V
⇐⇒ [x, y] = 0 on V
=⇒ dim V ≤ n
⇐⇒ V coisotropic
⇐= dim V = 1.
⇐⇒ V ∩ V = V 0 = V
⇐⇒ [x, y] = 0 on V
=⇒ dim V ≥ n
⇐⇒ V isotropic
⇐= dim V = 2n − 1.
2.3 Symplectic Space 139
(4) V Lagrangian : V = V
⇐⇒ V is isotropic and coisotropic
⇐⇒ V is isotropic and dim V = n
⇐⇒ V is coisotropic and dim V = n
⇐⇒ V maximally isotropic
⇐⇒ V minimally coisotropic.
(5) V non-degenerate : V ∩ V = {0}
⇐⇒ V + V = R2n
⇐⇒ [x, y] non-degenerate on V
⇐⇒ If [x, y] = 0 ∀ y ∈ V, then x = 0
⇐⇒ V non-degenerate
=⇒ dim V is even.
Proof. Without loss of generality, we can assume that dim U = k < n. Therefore,
there exists a vector ak+1 ∈ U \U = ∅, V = U + {ak+1 } ⊃ U, V is isotropic.
R2n = P1 "P
2"
· · · "P
n = U +̇V,
R2n = P1 "P
1 , P1 ∩ P1 = {0},
dim P1 = 2(n − 1), {a2 , a3 , · · · , an } ⊂ P1 is maximally isotropic in P1 , i.e.,
a Lagrangian subspace of P1 . By inductive assumption, there exists b2 , · · · , bn and
{b2 , · · · , bn } is Lagrangian space in P1 . Moreover,
Proof. Similarly manner, by induction with respect to dimension n, the proof can be
obtained.
U = U0 "
N, U 0 = U ∩ U .
Theorem 3.11. Let V1 , V2 be two disjoint isotropic subspaces. Then, there exist two
Lagrangian subspaces W1 , W2 that are disjoint. W1 ∩ W2 = {0} such that W1 ⊃
V1 , W2 ⊃ V2 .
Evidently,
Ui0 = Vi ∩ U 0 = Vi ∩ U ∩ U
= U ∩ U = U 0.
W ∩ U = {0}, U W = Rm .
Therefore,
Rα,α ∩ L = (Rα,α ∩ Rν,0 ) ∩ (Rν,0 ∩ L)
= Rα,0 ∩ (Rν,0 ∩ L) = {0}.
Then,
Rα,α + L = R2n .
The theorem is proved.
3. Matrix representation of subspaces in R2n
⎡ ⎤
a11 · · · a1k !
⎢ ⎥ A1
A = ⎣ ... ..
. ⎦ = [a1 , a2 , · · · , ak ] = ,
A2
a2n,1 · · · a2n,k
where A ∈ M (2n, k), ai ∈ M (2n, 1), Aj ∈ M (n, k).
Definition 3.15. A ∈ M (2n, k) is non-singular, if rankA = k.
Let A ∈ M (2n, k), B ∈ M (2n, l). Then, [A, B] ∈ M (2n, k + l). If [A, B] is
non-singular, then both A and B are non-singular.
G2n,k = {all k-dim subspaces in R2n }, called as Grassmann manifold.
If A ∈ M (2n, k) is non-singular, we define {A} = {a1 , · · · , ak } to be a k-dim
subspace in R2n generated by k column vectors a1 , · · · , ak of A.
144 2. Symplectic Algebra and Geometry Preliminaries
Let
2.3 Symplectic Space 145
!
A1
A= = [a1 , · · · , an ] ∈ M (2n, n),
A2
!
B1
B= = [b1 , · · · , bn ] ∈ M (2n, n),
B2
!
A1 B1
M = [A, B] = = [a1 , · · · , an , b1 , · · · , bn ] ∈ M (2n),
A2 B2
M ∈ Sp(2n) ⇐⇒ M JM = J
⇐⇒ A, B are a symmetric pair, A is conjugate to B
⇐⇒ A JA = On = B JB, A JB = In
⇐⇒ A1 A2 − A2 A1 = On = B1 B2 − B2 B1 , A1 B2 − A2 B1 = In
⇐⇒ ai Jaj = bi Jbj = 0, ai Jbj = δij ,
M ∈ CSp(2n) ⇐⇒ ∃μ = 0, M JM = μJ
⇐⇒ A, B form a symmetric pair and A is conformally conjugate to B
K −1 = (M −1 KM −1 )−1 = M K −1 M .
det [A, B] = 0.
Theorem 3.32. For two mutually transversal Lagrangian subspaces {A}, {B}, there
always exists a third Lagrangian subspace {C}, transversal to {A} and {B}.
Proof. Take {a1 , a2 , · · · , an } = {A}, {B} = {b1 , b2 , · · · , bn }, such that
A JA = O, B JB = O, A JB = In , B JA = −In .
Set C = A + B. Then,
(A + B) J(A + B) = (A + B )J(A + B)
= A JA + A JB + B JA + B JB = O;
and
det [A, A + B] = det [A, B] = 0,
det [B, A + B] = det [A, B] = 0.
The theorem is proved.
Theorem 3.33. For any two Lagrangian subspaces {A}, {B} there exists another
Lagrangian subspace {C}, transversal to {A} and {B}.
Proof. Assume U 0 = {A, B} = 0. Take
such that
{a1 , · · · , ak , c1 , · · · , cn−k } = {A},
{b̄1 , · · · , b̄k , c1 , · · · , cn−k } = {B},
[ai , aj ] = 0 = [b̄i , b̄j ],
[ai , b̄j ] = δij , i, j = 1, · · · , k.
=⇒ M = [A, J −1 A], A JA = On , A A = In .
!
A1 O
(I) Sp(2n) M = , which are diagonal blocks
O B1
!
A1 O
=⇒ M = , A1 ∈ GL(n, R).
O A−1
1
! !
I B1 I S
(II) Sp(2n) M = =⇒ M = , S = S.
O B2 O I
! !
A1 O I O
(II ) Sp(2n) M = =⇒ M = , S = S.
A2 I S I
!
Iα Iα
(III) Sp(2n) M = Jα = , α ⊂ ν, and symplectic substitution
−Iα Iα
!
Iα Iα
Jα = .
−Iα Iα
M = [A, B] ∈ Sp2n =⇒ A JA = O, A JB = I.
M = [A, B] ∈ O2n =⇒ A B = O, B B = I.
7 ⎫
A JA = O ⎪
=⇒ A (JA + B) = O ⎪
⎪
⎪
⎬ !
AB=O A
=⇒ (JA + B) = O
/ ⎪
⎪ B
B JA = −I ⎪
=⇒ B (JA + B) = O ⎪
⎭
BB = I
=⇒ M (JA + B) = O =⇒ JA + B = O =⇒ B = J −1 A.
Lemma 3.35. There exist polynomials ϕn (A) = ϕn (ajk ), ψn (A) = ψn (ajk ) in 2n×
n variables a11 , · · · , a2n,n with integer coefficients such that
150 2. Symplectic Algebra and Geometry Preliminaries
!
A1 −A2
det [A J A] = det = (ϕn (A))2 + (ψn (A))2 ≥ 0,
A2 A1
where ⎡ ⎤
a11 ··· a1n
⎢ .. ⎥ .
A = ⎣ ... . ⎦
a2n,1 · · · a2n,n
Proof.
A1 −A2 A + i A2 −A2 + i A1
= 1
A2 A1 A2 A1
A1 + i A2 −A2 + i A1 − i (A1 + i A2 )
=
A2 A1 − i A2
A1 + i A2 O
=
A2 A1 − i A2
= (A1 + i A2 ) (A1 − i A2 )
= (A1 + i A2 ) (A1 + i A2 )
= (Re|A1 + i A2 |)2 + (Im|A1 + i A2 |)2
= (ϕn (A))2 + (ψn (A))2 ≥ 0.
Therefore, the lemma is completed.
Theorem 3.36. M ∈ Sp2n (0) = Un ⇒ |M | = 1.
Proof. We will prove that
7
M M = I =⇒ |M |2 = 1 =⇒ |M | = ±1
=⇒ |M | = 1.
By Lemma 3.35, M = [A, J −1 A] =⇒ |M | ≥ 0
The proof can be obtained.
(2) Sp2n (I).
!
A1 O
M= , in which diagonal blocks ∈ GL(n, R),
O B1
!
A1 O
M= , A1 ∈ GL(n, R), |M | = |A1 ||A1 |−1 = 1,
O A−1
1
! ! !
A1 O T O P O
= ,
O A−11
O T O P −1
!
T O
∈ On ⊂ Un ,
O T
2.3 Symplectic Space 151
where
' ! /
T O
On = , T T = In = Un ∩ GL(n, R) = Sp2n (0) ∩ Sp2n (I),
O T
' ! /
P O
, P = P > 0 is not a group,
O P −1
' ! /
P O
, P = P > 0 ∩ Sp2n (0) = {I2n }.
O P −1
M ∈ Sp2n (I,II) =⇒ |M | = 1.
Sp2n (II) ∩ Sp2n (0) = Sp2n (II ) ∩ Sp2n (0) = {I}.
such that
M {A} = {B}.
{C} = {A}, C JC = O, C C = I.
{D} = {B}, D JD = O, D D = I.
MC , MD ∈ Sp2n (0) = Un .
Thus, M C = D.
We obtain M {C} = {M C} = {D}, i.e., M {A} = {B}.
Theorem 3.39. Let G = CSp(2n), Sp(2n), or Sp2n (0) = Un . Then, the following
action of G on Λn : M {A} = {M A}(A ∈ Λn , M ∈ G) is
1◦ Transitive, i.e., ∀{A}, {B} ∈ Λn , ∃M ∈ G, such that M {A} = {B}.
2◦ If for any {A} ∈ Λn , M {A} = {A} then M = ±I2n when G = Sp(2n) or
Sp2n (0); and M = μI2n when G = CSp(2n).
! ! ! ! !
A B O B O O
= = Q= ,
C D I D I Q
' /
I
and so B = C = O. Again, take {A} = , P = P . Then, from the equality
P
' / ' /
I I
M = ,
P P
it follows that
! ! ! ! !
A O I A I Q
= = Q= ,
O D P DP P PQ
AP = P A, ∀P = P.
It follows from the Theorem 3.38 that ϕ is surjective. We will prove that ϕ is injective.
' / ' / ' / ' /
I I −1 I I
If M1 = M2 , then M2 M1 = .
O O O O
! ' / ' /
A B I I
Set M = M2−1 M1 = ∈ Sp2n (0). Then, M = , i.e.,
C D O O
! ! ! !
A B I A I
= = Q.
C D O C O
154 2. Symplectic Algebra and Geometry Preliminaries
Thus, C = O.
!
A B
Since ∈ Sp2n (0), by Theorem 3.34:
C D
! ! ! ! !
B A O −I A O
= J −1 = =
D O I O O A
and !
A
[A , O] = A A = I.
O
!
A O
This means that M = , A A = I ∈ On .
O A
is a bijection.
Proof. If M1 Sp2n (I, II) = M2 Sp2n (I, II), then M1 ∈ M2 Sp2n (I, II), i.e.,
!
Q QS
M1 = M2 for some Q ∈ GL and S ∈ SM (n).
O Q−1
Thus,
' / ! ! !
I Q QS I Q
M1 = M2 P = M2 P
O O Q−1 O O
! ' /
I I
= M2 QP = M2 .
O O
Thus, C = O.
2.3 Symplectic Space 155
! ! !
A B A O I B1
M= = ∈ Sp(2n),
C D O A−1 O D1
!
−1 A O
where B1 = A B, D1 = A D. Since M ∈ Sp(2n) and ∈ Sp(2n), −1
O A
!
I B1
must be symplectic too.
O D1
!
I B1
By definition, ∈ Sp2n (II). Therefore, M ∈ Sp2n (I, II).
O D1
= M 0 M1 M2 ,
! !
Q O I C1
where M1 = −1 ∈ Sp2n (I), M2 = ∈ Sp2n (II).
O Q O C2
156 2. Symplectic Algebra and Geometry Preliminaries
where
!
T O
M0 = M 0 ∈ Sp2n (0), M2 = M 2 ∈ Sp2n (II),
O T
!
P O
M1 = , P = P > 0.
O P −1
Thus,
2.3 Symplectic Space 157
! !
P1 (S1 − S)P P1 P −1
= J
P1−1 P O
! !
O −I P1 P −1
=
I O O
!
O
= .
P1 P −1
Then, P1 (S1 − S)P = O and P1−1 P = P1 P −1 , i.e., S1 = S and P12 = P 2 .
−1
Therefore, P1 = P , since P1 , P are positive definite. It follows that, M01 M0 = I,
i.e., M0 = M01 . This is what we need to prove.
!
A1 B1
Theorem 3.45. If M = ∈ Sp(2n) with |A1 | = 0, then it can be de-
A2 B2
composed as
M = M 2 M1 M2 ,
where M2 ∈ Sp2n (II ), M1 ∈ Sp2n (I), M2 ∈ Sp2n (II).
This leads to |M | = 1.
Proof. By Gauss elimination,
! ! !
A1 B1 I −A−1 1 B1 A1 O
= ,
A2 B2 O I A2 B2 − A2 A−1
1 B1
! ! !
I O A1 O A1 O
= .
−A2 A−1
1 I A2 B2 − A2 A−1
1 B1 O B2 − A2 A−1 1 B1
Thus,
A−1
1 = B2 − A−1 −1 −1
1 A2 B1 = B2 − (A2 A1 ) B1 = B2 − A2 A1 B1 .
Obvioualy,
! !
A1 O A1 O
M1 = −1 = ∈ Sp2n (I),
O B2 − A2 A1 B1 O A−1
1
!−1
I O
M2 = ∈ Sp2n (II ),
−A2 A−1
1 I
!−1
I −A−1 1 B1
M2 = ∈ Sp2n (II),
O I
M = M3 M2 M1 M2 ,
where M3 ∈ Sp2n (III), M2 ∈ Sp2n (II ), M1 ∈ Sp2n (I), M2 ∈ Sp2n (II), and it
reduces |M | = 1 too.
!
A1 B1
Proof. It follows from that there exists α ⊂ ν such that Jα−1 =
A2 B2
!
C1 D1
with |C1 | = 0 and Theorem 3.45.
C2 D2
Definition 3.52. A real polynomial P2n (λ) = a0 λ2n + a1 λ2n−1 + · · · + a2n is even
if P (λ) = P (−λ).
2.3 Symplectic Space 159
(2)
(2) (2) (4) (2)
Obviously, P2n (λ) is even iff a2i+1 = 0 (i = 0, 1, · · · , n − 1). Every even poly-
nomial P2n (λ) can be rewritten in the following form
;
n
P2n (λ) = a0 (λ2 − ci ).
i=1
(2)
(4)
(2) (2) (2)
1
where we define ϕ(q) = q Sq, called as a generating function [Wei77,FWQW89,Fen95,Ge91]
2
for L.
Remark 3.55. There exists a generating function for Lagrangian subspace transversal
to Rν,0 or R0,ν .
Theorem 3.56. For a non-singular symmetric pair A, there is α ⊂ ν such that
det(Jα−1 A)2 = 0 and det(Jα−1 A)1 =
0, where Jα−1 A and Jα−1 A are non-singular
symmetric pairs.
Proof. By Theorem 3.14, there exists
! α ⊂ ν such that {A}+̇{Rα,α } = R2n . This
Iα A1
shows that the matrix is non-singular.
Iα A2
Multiplying by Jα−1 , we have
! ! ! !
−1 Iα A1 Iα −Iα Iα A1 O (Jα−1 A)1
Jα = = .
Iα A2 Iα Iα Iα A2 I (Jα−1 A)2
2.3 Symplectic Space 161
Elements of Sp(K, 2n) or CSp(K, 2n) are called as K-symplectic matrices and
conformally K-symplectic matrices respectively.
Theorem 3.59. Let M be non-singular.
1◦ M ∈ CSp(K, 2n), iff A ∈ Λn (K) ⇒ M A ∈ Λn (K).
2◦ M ∈ Sp(K, 2n), iff (M A) K(M B) = A KB or
M KM = K.
have the same power. In fact, ∃ Q1 , Q2 ∈ GL, such that K1 = Q1 JQ1 , K2 =
Q2 JQ2 ,
M K2 M = K1 ⇐⇒ M Q2 JQ2 M = Q1 JQ1 ,
i.e.,
Q−1 −1
1 M Q2 JQ2 M Q1 = J.
We have the following theorem similar to Theorem 3.23 and Theorem 3.59.
Theorem 3.62. Let M be non-singular. Then,
1◦ M ∈ CSp(K2 , K1 ), iff M A ∈ Λn (K1 ), ∀A ∈ Λn (K2 ).
2◦ M ∈ Sp(K2 , K1 ), iff (M A) K1 (M B) = A K2 B, ∀A, B ∈ Λn (K2 ).
Hamiltonian mechanics is geometry in phase space. Phase space has the structure of a
symplectic manifold.
ω(ξ, η) = 0, ∀η ∈ Tx M,
then ξ = 0 (nondegenerate).
The pair (M, ω) is called a symplectic manifold. We call it a presymplectic (almost
symplectic) manifold, if only condition 1◦ (2◦ ) is satisfied.
166 3. Hamiltonian Mechanics and Symplectic Geometry
Example 1.2. Consider the vector space R2n with coordinates (pi , qi ). Let ω =
n
d pi ∧ d qi . Then, ω defines a symplectic structure. Given two tangent vectors
i=1
we have
n
ω(ξ 1 , ξ 2 ) = ηi1 · ξi2 − ηi2 · ξi1 .
i=1
This example shows that any symplectic manifold can have standard symplectic
structure at least locally.
Exercise 1.3. Verify that (R2n , ω) is a symplectic manifold. For n = 1, ω becomes
an area measure on a plane.
n
∂ ∂
n
∂
ξ= ai + bi , π∗ (ξ) = bi .
i=1
∂pi ∂qi i=1
∂qi
n
Therefore, p(π∗ (ξ)) = pi bi , which results in
i=1
n
n
σ= pi dqi , ω = dσ = dpi ∧ dqi .
i=1 i=1
3.1 Symplectic Manifold 167
which has an expression under local basis for the tangent field
∂ ∂
Hp − Hq .
∂q ∂p
168 3. Hamiltonian Mechanics and Symplectic Geometry
and
n
ω|u = d xi ∧ d yi .
i=1
n
Proof. The sufficiency can be easily derived, since dxi ∧ dyi is a closed form.
i=1
Necessity. We first assume that M = E is a linear space, and m = 0 ∈ E.
Let ω1 be the constant-form ω(0), ω " = ω1 − ω, ωt = ω + t" ω (0 ≤ t ≤ 1).
For each t, ωt (0) = ω(0) is nondegenerate. Hence, by the openness of the set of
isomorphisms from E to E ∗ , there exists a neighborhood of 0 on which ωt (0 ≤
t ≤ 1) is nondegenerate for all 0 ≤ t ≤ 1. We can assume that this neighborhood
" = dθ. Without loss
is a ball. Thus by Poincaré lemma, there exists a 1-form θ s.t. ω
of generality, we assume θ(0) = 0. Since ωt is nondegenerate, there exists a smooth
vector field X, s. t. iX ωt = −θ. Since Xt (0) = 0, from the local existence theory
of ODEs, there is a sufficiently small ball on which the integral curves of Xt are
well defined for t ∈ [0, 1]. Let Ft be the flow starting at F0 = identity. By the Lie
derivative formula for a time-dependent vector field, we have
d d
(F ∗ ωt ) = Ft∗ (LXt ωt ) + Ft∗ ωt
dt t dt
= Ft∗ d iXt ωt + Ft∗ ω
" = Ft∗ (−d θ + ω
" ) = 0.
Definition 2.1. A fundamental differential 1-form and 2-form in R2n are defined by
the following formulae:
n
n
1-form : θ = pi d qi = zi d zn+i ;
i=1 i=1
n
n
2-form : ω = d θ = d p ∧ d q = d pi ∧ d qi = d zi ∧ d zn+i .
i=1 i=1
n
n
ω(ξ, η) = d zi ∧ d zn+i (ξ, η) = (ξi ηn+i − ηi ξn+i )
i=1 i=1
⎡ ⎤
5 6 η1
O In ⎢ .. ⎥
= (ξ1 , · · · , ξ2n ) ⎣ . ⎦
−In O
η2n
= ξ Jη, (2.1)
!
O In
where J = , and (ξ1 , · · · , ξ2n ) represents the vector ξ ∈ Tz R2n con-
−In O
sisting of components ξi . The Equation (2.1) is the matrix representation of the 2-form
ω on Tz R2n .
i.e.,
ωξ1 (η) = Ωξ(η) = ω 2 (η, ξ) = (−iξ ω)η, ∀ η ∈ Tz R2n ,
or
ωξ1 = Ωξ = −iξ ω.
From the equation ω(η, ξ) = η Jξ = (Jξ) η, it follows that
ωξ1 = Ωξ = Jξ
or
2n
ωξ1 = Ωξ = (Jξ)i dzi
i=1
Obviously, Ω is an isomorphism from the tangent space Tz R2n into the cotangent
space Tz∗ R2n . This naturally induces a mapping from X (R2n ) into Ω1 (R2n ):
1
ωX = Ω(X)(z) = Ω(X(z)) = −iX ω, ∀ X ∈ X (R2n ).
XH = J −1 Hz ,
d H = ΩXH = −iX ω.
3. Canonical systems
Now, we consider a canonical equation in R2n .
Definition 2.3.
dz
= J −1 Hz (2.2)
dt
or
dp dq
= −Hq , = Hp . (2.3)
dt dt
3.2 Hamiltonian Mechanics on R2n 171
The phase flow of the Hamiltonian vector field is denoted as φtH , and called the
Hamiltonian phase flow.
(φtH )∗ ω = ω.
Proof. Since
d t ∗ d
(φ ) ω = (φt+s )∗ s=0 ω
dt H ds H
d
= (φtH )∗ · (φsH )∗ s=0 ω
ds
= (φtH )∗ LXH ω,
and
LXH ω = (iXH d + d iXH )ω = iXH d ω + d iXH ω
= 0 + (−d (dH)) = 0,
we have
d t ∗
(φ ) ω = 0,
dt H
i.e.,
(φtH )∗ ω = (φtH )∗ t=0 ω = ω.
The theorem is proved.
4. Integral invariants[Arn89]
Let g : R2n → R2n be a differentiable mapping.
g∗ ωk = ωk .
172 3. Hamiltonian Mechanics and Symplectic Geometry
Theorem 2.8. If the forms ω k and ω l are integral invariants of the map g, then the
form ω k ∧ ω l is also an integral invariant of g.
the integral of ω 2k is equal to the sum of the oriented volume of projections onto
the coordinate planes (pi1 , · · · , pik , qi1 , · · · , qik ). Therefore, a Hamiltonian phase flow
preserves the sum of the oriented area as projections onto the coordinate planes
(pi1 , · · · , pik , qi1 , · · · , qik ) (1 ≤ k ≤ n).
(g ∗ ω)(ξ, η) = ω(g∗ ξ, g∗ η) = ξ M JM η,
∂g
where M = g∗ = is Jacobian of g.
∂z
3.2 Hamiltonian Mechanics on R2n 173
g canonical ⇐⇒ M JM = J, ∀ z ∈ R2n ,
⇐⇒ ξ M JM η = ξ Jη, ∀ ξ, η ∈ Tz R2n , z ∈ R2n ,
⇐⇒ g ∗ ω(ξ, η) = ω(ξ, η), ∀ ξ, η ∈ Tz R2n ,
⇐⇒ g ∗ ω = ω.
M JM = μJ,
Proof.
d z ∂ z d z ∂g dz ∂g
= + =M + .
dt ∂ z dt ∂t dt ∂t
Set H(z) = H(g(z, ◦ g; then H z = ∂ z H
t), t) = H z . Thus, from the equation
∂z
d z z ,
= J −1 H
dt
we have
dz ∂g
M + = J −1 M −1 H z ,
dt ∂t
174 3. Hamiltonian Mechanics and Symplectic Geometry
i.e.,
dz ∂g
= M −1 J −1 M −1 H z −
dt ∂t
∂g
= J −1 JM −1 J −1 M −1 H z − JM −1
∂t
= J −1 (u + v),
∂g
where u = BH z , B = JM −1 J −1 M −1 , v = C , C = −JM −1 , and u depends
∂t
as well as on z, and v depends only on z.
on the Hamiltonian H
∈ C ∞ (R2n ), there exists another function H(z), such that
For every H
dz
= J −1 Hz ,
dt
as a constant, we get
In the above equation , taking H
∂ vi ∂ vk
= , i, k = 1, · · · , 2n. (2.5)
∂ zk ∂ zi
Consequently,
∂ ui ∂ uk z ),
= , ∀ H( i, k = 1, · · · , 2n. (2.6)
∂ zk ∂ zi
2n
Notice that ui = (BH z )i = Bij H zj , (2.6) becomes
j=1
∂
2n 2n
∂H ∂ ∂H
Bij = Bkj . (2.7)
∂ zk j=1 ∂ zj ∂ zi ∂ zj
j=1
Take H(
z ) = zl ◦ g −1 (l = 1, · · · , 2n), then H(z) = zl (l = 1, · · · , 2n). By this,
Equation (2.8) gets split into classes of equations:
3.2 Hamiltonian Mechanics on R2n 175
∂ Bij ∂ Bkj
= , i, k, j = 1, · · · , 2n, (2.9)
∂ zk ∂ zi
2n
2n
∂2 H ∂2 H
Bij = Bkj , i, k = 1, · · · , 2n. (2.10)
∂ zk ∂ zj ∂ zi ∂ zj
j=1 j=1
∂2 H
Set A = . Obviously, A is symmetric, i.e., A = A. Then, (2.10) indicates
∂ zk ∂ zj
BA = (BA) = A B = AB , ∀ A = A.
This implies
B = μ(z, t)I, (2.11)
where μ = 0. Since |B| = 0, or
we have
∂ gj ∂ z ∂g
vi = Cij = −μ(t) l Jlj j ,
∂t ∂ zi ∂t
∂ vi ∂ ∂ zl ∂g
= − μ(t) Jlj j
∂ zk ∂ zk ∂ zi ∂t
2
∂ zl ∂g ∂ z ∂ 2 gj
= −μ(t) Jlj j + l Jlj .
∂ zk ∂ zi ∂t ∂ zi ∂ zk ∂ t
∂ 2 zl ∂g ∂ z ∂ 2 gj
−μ(t) Jlj j + l Jlj
∂ zk ∂ zi ∂t ∂ zi ∂ zk ∂ t
2
∂ zl ∂g ∂ zl ∂ 2 gj
= μ(t) Jlj j + Jlj ,
∂ zi ∂ zk ∂t ∂ zk ∂ zi ∂ t
i.e.,
∂zl ∂ zj ∂ zl ∂ zj
Jlj = Jlj ,
∂ zi ∂ zk t ∂ zk ∂ zi t
i.e.,
(M JMt )ik = (M JMt )ki ,
which shows show that M JMt is a symmetric matrix. Therefore,
Then, we have
= −(φp ψq − φq ψp ).
Property 2.17. Let φ, ψ, χ be smooth functions on R2n , then the Poisson bracket has
following basic properties:
1◦ anti-symmetric: {φ, ψ} = −{ψ, φ}.
2◦ bilinear: {αφ + βψ, χ} = α{φ, χ} + β{ψ, χ}, α, β ∈ R.
3◦ Jacobi identity: {{φ, ψ}, χ} + {{ψ, χ}, φ} + {{χ, φ}, ψ} = 0.
1◦ and 2◦ are self-evident. The Jacobi identity can be proved by direct computa-
tion, but it also follows from the following proposition and the corresponding Jacobi
identity of the vector field.
Proof. By definition,
2n
where Xφ = Ω−1 d φ = (J −1 φz )i d zi . The Equations (2.13) and (2.14) are just
i=1
1◦ and 2◦ of Proposition 2.18 respectively. However,
= {φ, ψ}P,Q .
Theorem 2.20. A function F is a first integral of the phase flow with the Hamiltonian
H iff its Poisson bracket with H is identically zero:
{F, H} = 0.
Proof. By the 4◦ of proposition above,
d
LXH F = (φtH )∗ F = {F, H} = 0.
dt t=0
Thus,
d d t ∗ d
F (φtH (z)) = (φ ) F (z) = (φt+s )∗ F (z)
dt dt H d s s=0 H
∗
d ∗
= (φt φs ) F (z)
d s H H s=0
∗ d ∗
= φtH (φs )∗ F (z) = φtH LXH F (z) = 0,
d s H s=0
i.e., F is a first integral of the phase flow with the Hamiltonian H. The necessary
condition is evident.
3.2 Hamiltonian Mechanics on R2n 179
Theorem 2.21. H is a first integral of the phase flow with Hamiltonian function H.
This follows immediately from the Theorem 2.21 and the fact that {F, H} =
−{H, F }.
Theorem 2.23 (Poisson theorem). The Poisson bracket of the two first integrals
F1 , F2 of a system with a Hamiltonian function H is again a first integral.
Definition 2.25. A linear subspace of a Lie algebra is called a subalgebra if the sub-
space is closed under the commutator, i.e., the commutator of any two elements of the
subspace belongs to it.
Evidently, a subalgebra of a Lie algebra is itself a Lie algebra with the original
commutator.
By the proposition and theorems above, we have:
Corollary 2.26. The Hamiltonian vector fields on R2n form a subalgebra of the Lie
algebra of all vector fields.
Corollary 2.27. The commutator of the Hamiltonian phase flow with a Hamiltonian
form a subalgebra of the Lie algebra of all functions.
180 3. Hamiltonian Mechanics and Symplectic Geometry
⎡ ∂P ⎤ ⎡ % &−1 ⎤ ⎡
∂P ∂Q ∂P ∂X ⎤ ⎡ ∂ ⎤
⎢ ∂x ⎥ ⎢ ∂x ∂q ⎥ ⎣ ∂x ∂q ⎦ ⎣ ∂q (P ◦ X)
Tz S = ⎣ ⎦=⎣ ⎦= = ⎦.
∂Q
I I I
∂x
Setting f = P ◦ X, we get
5 6 7 7 7 ⎡ ⎤
∂f
P (x) P ◦ X(q) f (q)
S= ∈ R2n = = , Tz S = ⎣ ∂q ⎦,
Q(x) q q I
∂f
If f (q) is a function with Jacobian M = , and S = Gf , the graph of f , then
∂q
5 6 7
f (q)
W (S) = z = W (z) | z = ∈S
q
d z(t)
= J −1 Hz , z(0) = z0 , (2.15)
dt
with the Hamiltonian H(z) = H(p, q). Let z(t) = (p(t), q(t)) be its solution and
G(t) the 1-parameter group of diffeomorphisms in R2n .
5 6 5 6
p0 p(t)
G(t) : z0 = −→ z(t) = G(t)z0 = , G(0) = I.
q0 q(t)
p
0
Let M0 = be an n-dim initial manifold, p0 a function of q0 , and M0
q0
p
0 (q0 ) ∂ p0
form a Lagrangian manifold, i.e., M0 = and ∈ Sm. Since G(t) :
q0 ∂ q0
p p(t)
0
→ is a canonical transformation for a fixed t in some neighbourhood
q0 q(t)
2n
of R , and
3.2 Hamiltonian Mechanics on R2n 183
⎡ ⎤
5 6 ∂p ∂p
A B ⎢ ∂ p0 ∂ q0 ⎥
G∗ (t) = =⎢ ⎥
C D
⎣ ∂q ∂q ⎦
∂ p0 ∂ q0
is a symplectic matrix,
⎡ ⎤
5 6 ⎡ ∂ p0 ⎤ ∂p
A 0 +B
Y1 ⎢ ∂ q0 ⎥
= G∗ ⎣ ∂ q0 ⎦ = ⎢
⎣ ∂ p0
⎥
⎦
Y2 I C +D
∂ q0
Y N
∂p 1
is a symmetric pair. If C 0 + D = 0, then ∼ , where
∂q0 Y2 I
−1
∂p ∂p
N = A 0 +B C 0 +D ∈ Sm(n).
∂ q0 ∂ q0
ω 1 = p d q + H d t.
ω 1 = p d q + H d t = d ϕ,
∂p
is symmetric. We know that the matrix is symmetric, and so we need to only prove
∂q
∂p ∂H
= .
∂t ∂q
In addition,
184 3. Hamiltonian Mechanics and Symplectic Geometry
∂H ∂ H(p(q, t), q) ∂ p
=− =− H p − Hq .
∂q ∂q ∂t
Thus,
∂p ∂p ∂H
= −Hq − Hp = .
∂t ∂q ∂q
Consequently, there exists a scalar function ϕ(q, t), s.t.
p d q + H d t = d ϕ.
or
ϕt + H(ϕq , q) = 0,
which is called the Hamilton–Jacobi equation.
Bibliography
[AA88] D.V. Anosov and V.I. Arnold: Dynamical Systems I. Springer, Berlin, (1988).
[AA89] V. I. Arnold and A. Avez: Ergodic Problems of Classical Mechanics. Addison-Wesley,
New York, (1989).
[Abd02] S. S. Abdullaev: The Hamilton-Jacobi method and Hamiltonian maps. J. Phys. A:
Math. Gen., 35(12):2811–2832, (2002).
[AKN78] V. I. Arnold, V. V. Kozlov, and A. I. Neishtadt: Mathematical Aspects of Classical
and Celestial Mechanics. Springer, Berlin, Second edition, (1978).
[AM78] R. Abraham and J. E. Marsden: Foundations of Mechanics. Reading, MA: Addison-
Wesley, Second edition, (1978).
[AMR88] R. Abraham, J. E. Marsden, and T. Ratiu: Manifolds, Tensor Analysis, and Applica-
tions. AMS 75. Springer-Verlag, Berlin, Second edition, (1988).
[AN90] A. I. Arnold and S.P. Novikov: Dynamical System IV. Springer Verlag, Berlin, (1990).
[Arn88] V. I. Arnold: Geometrical Methods in The Theory of Ordinary Differential Equations.
Springer-Verlag, Berlin, (1988).
[Arn89] V. I. Arnold: Mathematical Methods of Classical Mechanics. Berlin Heidelberg:
Springer-Verlag, GTM 60, Second edition, (1989).
[Ber00] R. Berndt: An Introduction to Symplectic Geometry. AMS Providence, Rhode Island,
(2000).
[Bir23] G. D. Birkhoff: Relativity and Modern Physics. Harvard Univ. Press, Cambridge,
Mass., Second edition, (1923).
[BK89] G.W. Bluman and S. Kumei: Symmetries and differential equations. AMS 81.
Springer-Verlag, New York, (1989).
[Car65] C. Carathe’odory: Calculus of Variation and Partial Differential Equations of First
Order, Vol.1. Holden-Day, San Franscisco, (1965).
[Car70] H. Cartan: Differential Forms. Houghton-Mifflin, Boston, (1970).
[CH53] R. Courant and D. Hilbert: Methods of Mathematical Physics. Interscience, New York,
Second edition, (1953).
[Che53] S. S. Chern: Differential Manifolds. University of Chicago, (1953). Lecture notes.
[Fen86] K. Feng: Difference schemes for Hamiltonian formalism and symplectic geometry. J.
Comput. Math., 4:279–289, (1986).
[Fla] H. Flanders: Differential Forms. Academie Press, New York, Second edition, (1963).
[FQ87] K. Feng and M.Z. Qin: The symplectic methods for the computation of Hamiltonian
equations. In Y. L. Zhu and B. Y. Guo, editors, Numerical Methods for Partial Differential
Equations, Lecture Notes in Mathematics 1297, pages 1–37. Berlin, Springer, (1987).
[FQ91a] K. Feng and M.Z. Qin: Hamiltonian Algorithms for Hamiltonian Dynamical Systems.
Progr. Natur. Sci., 1(2):105–116, (1991).
[FQ91b] K. Feng and M.Z. Qin: Hamiltonian algorithms for Hamiltonian systems and a com-
parative numerical study. Comput. Phys. Comm., 65:173–187, (1991).
[FQ03] K. Feng and M. Z. Qin: Symplectic Algorithms for Hamiltonian Systems. Zhejiang
Science and Technology Publishing House, Hangzhou, in Chinese, First edition, (2003).
186 Bibliography
[FWQW89] K. Feng, H. M. Wu, M.Z. Qin, and D.L. Wang: Construction of canonical dif-
ference schemes for Hamiltonian formalism via generating functions. J. Comput. Math.,
7:71–96, (1989).
[Gol80] H. Goldstein: Classical Mechanics. Addison-Wesley Reading, Massachusetts, (1980).
[GS84] V. Guillemin and S. Sternberg: Symplectic Techniques in Physics. Cambridge Univer-
sity Press, Cambridge, (1984).
[Lan95] S. Lang: Differential and Riemannian Manifolds. Springer-Verlag, Berlin, (1995).
[LL99] L. D. Landau and E. M. Lifshitz: Mechanics, Volume I of Course of Theoretical
Physics. Corp. Butterworth, Heinemann, New York, Third edition, (1999).
[LM87] P. Libermann and C.M. Marle: Symplectic Geometry and Analytical Mechanics. Rei-
del Pub. Company, Boston, First edition, (1987).
[Mac70] S. MacLanc: Hamiltonian mechanics and geometry. Amer. Math. Mon., 77(6):570–
586, (1970).
[Sie43] C.L. Siegel: Symplectic geometry. Amer. J. Math, 65:1–86, (1943).
[Tre75] F. Treves: Pseodo-Differential Operator. N.Y.: Acad. Press, First edition, (1975).
[Wei77] A. Weinstein: Lectures on symplectic manifolds. In CBMS Regional Conference, 29.
American Mathematical Society, Providence, RI , (1977).
[Wes81] C. Von. Westenholz: Differential Forms in Mathematical Physics. North-Holland,
Amsterdam, Second edition, (1981).
Chapter 4.
Symplectic Difference Schemes for
Hamiltonian Systems
The canonicity of the phase flow for time-independent Hamiltonian systems is one of
the most important properties. It ensures the preservation of phase areas and the phase
volume. Thus, preserving the canonicity of transition of difference schemes from one
time step to the next is also important in the numerical solutions of Hamiltonian sys-
tems. The goal of this chapter is to find some simple symplectic schemes, i.e., to
identify which one, among the existing difference schemes, is symplectic.
4.1 Background
It is well known that Hamiltonian systems have many intrinsic properties: the preser-
vation of phase areas of even dimension and the phase volume, the conservation laws
of energy and momentum, and other symmetries.
ṗ = −Hq , q̇ = Hp , (1.1)
!
p
where p = (p1 , · · · , pn )T , q = (q1 , · · · , qn )T . Let z = , and the standard
q
symplectic matrix be: 5 6
O In
J= , (1.2)
−In O
where In is the n × n identity matrix, and J has property J −1 = J = −J. Then,
system (1.1) can be written in a compact form:
ż = J −1 Hz , (1.3)
!
Hq
where Hz = ; H is called the Hamiltonian function of the system. The phase
Hp
t
flow of system (1.1) can be represented as gH . According to the fundamental theorem
188 4. Symplectic Difference Schemes for Hamiltonian Systems
for each pair of tangent vector ξ, η at point z ∈ Tz R2n , where J is the standard
symplectic structure Equation (??).
Let w : R2n → R2n be a differential mapping, z ∈ R2n → w(z) ∈ R2n ; the
corresponding Jacobian matrix is denoted by
⎡ ∂w ∂ w1 ⎤
1
···
⎢ ∂ z1 ∂ z2n ⎥
⎢ ⎥
⎢ .. .. ⎥
∂w ⎢ ⎥
=⎢ . . ⎥.
∂z ⎢ ⎥
⎢ ∂w ⎥
⎣ 2n
···
∂ w2n ⎦
∂ z1 ∂ z2n
The mapping w induces, for each z ∈ R2n , a linear mapping w∗ (z) from the tangent
space at z into the the tangent space at w(z):
∂w
ξ = (ξ1 , · · · , ξ2n ) −→ w∗ ξ = ξ.
∂z
! !
∂w ∂w
J = J, (1.5)
∂z ∂z
∂w
i.e., Jacobian is a symplectic matrix for every z.
∂z
Its geometric meaning is depicted in Fig. 1.1. According to the general theory of ODE,
for each Hamiltonian system (1.1), there corresponds a one parameter group of dif-
feomophisms g t , at least locally in t and z, of R2n such that
g 0 = id, g t1 +t2 = g t1 · g t2 .
q q
6 6 ∂w
∂z
ξ
[ ∂∂ w
z
]
ξ -
1 :
∂w
∂z
η
η
- p - p
O O
Fig. 1.1. Geometric meaning of preserving symplectic structure
z(t) = g t z(0).
- ··· - ···
ω J ∧ · · · ∧ ωJ = ωJ ∧ · · · ∧ ωJ , every 2n-chain σ 2n ⊂ R2n ,
g t σ 2n σ 2n
where the last one is the Liouville’s phase-volume conservation law. Another class of
conservation law is related to the energy and all the first integrals. A smooth function
ϕ(x) is said to be a first integral if ϕ(g t z) = ϕ(z), for all t, z. The latter is equivalent
to the condition {ϕ, H} = 0; H usually represents the energy, which is a first integral
of itself.
190 4. Symplectic Difference Schemes for Hamiltonian Systems
i.e.,
1
ωK (ξ, η)z = ξ K(z)η, K (z) = −K(z), det K(z) = 0.
2
A differentiable mapping w : R2n → R2n called K-symplectic, if w∗ ωK = ωK ,
i.e.,
% &
∂w ∂ w
K w(z) = K(z). (1.7)
∂z ∂z
The Darboux theorem establishes the equivalence between all symplectic struc-
tures. Every non-singular closed 2-form ωK can be brought to the standard form
Kij (z) d zi ∧ d zj = d ωi ∧ d wn+j
i<j i<j
JB + B J = O. (1.9)
All infinitesimal symplectic matrices form a Lie algebra with commutation operation
[A, B] = AB − BA, denoted as sp(2n). sp(2n) is the Lie algebra of the Lie group
Sp(2n). We have the following well-known proposition[FWQ90] , which can be found
in Chapter 2. Here, we omit the proof.
!
A B
Proposition 1.6. Let S = , A, B, C, D be an n × n matrix; then S ∈
C D
Sp(2n) iff:
S = M −1 N ∈ Sp(2n), iff M JM = N JN .
z n+1 − z n z n+1 + z n
=B . (2.7)
τ 2
The transition z n → z n+1 is given by
4.2 Symplectic Schemes for Linear Hamiltonian Systems 193
τ 1−λ
z n+1 = Fτ z n , Fτ = φ − B , φ(λ) = , (2.8)
2 1+λ
τ
where Fτ is a Cayley transformation of the infinitesimal symplectic − B, and is
2
symplectic according to Proposition 1.12.
The second scheme we consider is the staggered explicit scheme for a separable
Hamiltonian. For a separable Hamiltonian H(p, q) = U (p) + V (q),
p 1
1 1
H(p, q) = [p , q ] S = p U p + q V q = U (p) + V (q), (2.9)
2 q 2 2
where 5 6
U O
S= .
O V
U = U is positive definite and V = V , the canonical Equation (1.1), becomes:
dp dq
= −Vq , = Up . (2.10)
dt dt
The staggered explicit scheme is:
1 n+1 n+ 1
(p − pn ) = −Vq 2 , (2.11)
τ
1 1 1
q n+ 2 +1 − q n+ 2 = Upn+1 . (2.12)
τ
1
pT s are defined at integer time t = nτ , and q T s at half-integer times t = n + τ.
2
The transition 5 6 5 6
pn pn+1
n
w = 1
−→ 1
= wn+1
q n+ 2 q n+ 2 +1
is given by the following:
wn+1 = Fτ wn ,
where 5 6−1 5 6
I O I −τ V
Fτ = , (2.13)
−τ U I O I
as the product of two symplectic matrices, is symplectic (Proposition 1.7 ), and the
scheme has second order of accuracy.
nlm (x)
exp (x) ∼ = glm (x), (2.14)
dlm (x)
where
m
(l + m − k) ! m !
nlm (x) = xk , (2.15)
(l + m) ! k ! (m − k) !
k=0
l
(l + m − k) ! l !
dlm (x) = (−x)k . (2.16)
(l + m) ! k ! (l − k) !
k=0
For each pair of nonnegative integers l and m, the Taylor series expansion of
nlm (x)
about the origin point is:
dlm (x)
nlm (x)
exp (x) − = o(|x|m+l+1 ), |x| −→ 0, (2.17)
dlm (x)
and the resulting (l + m)-th order Padé approximation of exp (x) is denoted by glm .
Theorem 2.2. Let B be an infinitesimal symplectic; then, for sufficiently small |t|,
glm (tB) is symplectic iff l = m, i.e., gll (x) is the (l, l) diagonal Padé approximant to
exp (x).
Proof. Sufficiency. Let nll (x) = f (x) + g(x), dll = f (x) − g(x), where f (x) is an
even polynomial, g(x) is an odd one. In order to prove gtt (tB) ∈ Sp(2n), we only
need to verify Proposition 1.9 .
f (tB)+g(tB) J f (tB)+g(tB) = f (tB)−g(tB) J f (tB)−g(tB) . (2.18)
Comparing Equations (2.19) and (2.20) completes the proof of the “if” part of the
theorem.
The “only if” part. Without loss of generality, we may take l > m. We only need to
notice that in Equation (2.18), the order of the polynomial on the right hand is higher
than that on the left hand.
From Theorem 2.2, we can obtain a sequence of symplectic difference schemes based
on the diagonal (k, k) Padé table. In Table 2.1, the element of l-th row, m-th column is
denoted by (l, m). For the (1,1) approximation (i.e., l = 1, m = 1), we have the Euler
centered scheme
4.2 Symplectic Schemes for Linear Hamiltonian Systems 195
τB n
z n+1 = z n + (z + z n+1 ), (2.21)
2
Fτ(1,1) = φ(1,1) (τ B),
λ
1+
(1,1) 2.
φ (λ) = (2.22)
λ
1−
2
This scheme has second order accuracy.
For the (2,2) Padé approximation, we have:
τB n τ 2B2 n
z n+1 = z n + (z + z n+1 ) + (z − z n+1 ), (2.23)
2 12
whose transition is
(2,2)
Fτ = φ(2,2) (τ B),
λ λ2
1+
+
φ(2,2) (λ) = 2 12 . (2.24)
λ λ2
1− +
2 12
l
0 1 2 3 4
m
1 1 1 1 1
0
1 1−x x2 x2 x3 x2 x3 x4
1−x+ 1−x+ − 1−x+ − +
2 2 6 2 6 24
x x x x
1+x 1+ 1+ 1+ 1+
1 2 3 4 5
1 x 2x x3 3x x2 x3 4x 3x2 x3 x4
1− 1− + 1− + − 1− + − +
2 3 6 4 4 24 5 10 15 120 .
x2 2x x2 x x2 2x x2 x x2
1+x+ 1+ + 1+ + 1+ + 1+
+
2 2 3 6 2 12 5 20 3 30
1 x x x2 3x x2 x3 2x x2 x3 x4
1− 1− + 1− + − 1− + − +
3 2 12 5 20 60 3 20 30 360
4. Symplectic Difference Schemes for Hamiltonian Systems
x2 x3 3x x2 x3 3x 3x2 x3 x x2 x3 3x x2 x3
1+x+ + 1+ + + 1+ + + 1+ + + 1+ + +
3 2 6 4 4 24 5 20 60 2 10 120 7 14 210
1 x 2x x2 x x2 x3 4x x2 4x3 x4
1− 1− + 1− + − 1− + − +
4 5 20 2 10 120 7 7 210 840
det (I + B) = 0. (2.29)
whose inversion is
I + B = 2(I + S)−1 . (2.31)
[FWQ90]
Therefore S is non-exceptional, and we have the Cayley transformation :
and
B = (I − S)(I + S)−1 = (I + S)−1 (I − S). (2.33)
Let A be an arbitrary matrix. The equation
S AS = A (2.34)
expresses the condition that the substitution of S into both variables z, w leaves in-
variant the bilinear form z Aw.
Lemma 2.5. [Wey39] If the non-exceptional matrices B and S are connected by (2.32)
and (2.33), and A is an arbitrary matrix, then
S AS = A (2.35)
iff
B A + AB = O. (2.36)
B (I + S ) = I − S .
−AB = B A.
S (I + B ) = I − B
198 4. Symplectic Difference Schemes for Hamiltonian Systems
iff
B A + AB = O.
1
In other words, a quadratic form F (z) = z Az is invariant under the symplectic
2
transformation φ(τ B) iff F (z) is an invariant integral of the Hamiltonian system (2.1).
B A + AB = O.
for all τ with |τ | sufficiently small. Then, differentiating both sides of the above equa-
tion with respect to τ , we get
B ψλ (τ B) Aψ(τ B) + ψ(τ B) ABψλ (τ B) = O.
Setting τ = 0, it becomes
(B A + AB)ψ(0) ψλ (0) = O.
hold good for any analytic function ψ. From condition 2◦ , it follows that
Therefore,
d
ψ(τ B) A ψ(τ B)
dτ
d
= ψ(τ B ) A ψ(τ B)
dτ
= B ψλ (τ B ) Aψ(τ B) + ψ(τ B ) ABψλ (τ B)
= B Aψλ (−τ B) ψ(τ B) + ABψλ (−τ B) ψ(τ B)
= (B A + AB) ψλ (−τ B) ψ(τ B) = O,
i.e.,
ψ(τ B ) Aψ(τ B) = ψ(0) Aψ(0) = Aψ 2 (0) = A.
The proof is completed.
By taking successively A = J and A = A in Theorem 2.8 and using (2.3) – (2.6),
we obtain the following theorems.
Theorem 2.8. Take |τ | sufficiently small so that τ B has no eigenvalue at the pole
of the function φ(λ) in Theorem 2.7. Then, ψ(τ B) ∈ Sp(2n) iff B ∈ sp(2n). Let
B = J −1 C, C = C, A = A; then,
iff
AJC = CJA.
1
In other words, a quadratic form F (z) = z Az, is invariant under the symplectic
2
transformation ψ(τ B), iff F (z) is an invariant integral of the system (2.1).
The transformation φ(τ B) based on Theorem 2.7 includes exponential transfor-
mation exp(τ B), Cayley transformation ψ(−τ B/2), and diagonal Padé transforma-
tion as special cases. Taking φ(λ) in Theorem 2.7 as a rational function, then nec-
P (λ)
essarily ψ(λ) = , P (λ) is a polynomial, and is often normalized by setting
P (−λ)
P (0) = 1, P (0) = 0.
Theorem 2.9. Let P (λ) be a polynomial P (0) = 1, P (0) = 0, and
P (λ)
exp (λ) − = O (|λ|2k+1 ). (2.39)
P (−λ)
200 4. Symplectic Difference Schemes for Hamiltonian Systems
Then,
P (−τ B)z m+1 = P (τ B)z m ,
i.e.,
P (τ B) m
z m+1 = z (2.40)
P (−τ B)
is a symplectic scheme of order 2k for a linear system (2.1). This difference scheme
and the original system (2.1) have the same set of quadratic invariants.
P (x)
In order to find the approximate to exp (x) , we may express exp(x) in various
P (−x)
rational fraction ways. The following are examples:
nll (x) d (−x) 1 1
(1) exp (x) ∼ = ll . (2) exp (x) ∼ glm (x) · gml (x).
nll (−x) dll (x) 2 2
x x
1 + tanh e2
(3) exp (x) = 2 exp (x) =
x. (4) x .
e− 2
1 − tanh
2
1 x
(1 + e )
(5) exp (x) = 12 .
(1 + e−x )
2
Each denominator and numerator in the above expressions can be expanded about
the origin
in xTaylor
< series. The first term of the approximation gives the function
x
ψ(x) = 1 + 1 − , which yields the Euler centered scheme. Keeping m(> 1)
2 2
terms in the expansions for both the denominator and numerator, we will get a function
ψ(x) that will extend the Euler centered schemes. The schemes obtained in this way
are all symplectic; however, the order of accuracy of the first and third schemes is
higher than that of the last two kinds. For example, if in the formula (5) the first three
terms of the expansions of the denominator and numerator are retained, then the 4-th
order symplectic scheme is obtained. However, the same kind of truncation gives 6-th
order schemes from (1) and (3).
Proof.
= >
z k+1 − z k
B(z k+1 + z k ),
τ
= >
z k+1 + z k
= B(z k+1 + z k ), J −1 Hz
2
= k+1 >
z + zk
= (z k+1 + z k ), BJ −1 Hz = 0,
2
and so
Bz k , z k = Bz k+1 , z k+1 .
The proof is proved.
The last equation comes from the conservation law of original system.
Remark 3.2. As Euler centered schemes, high-order schemes constructed by the di-
agonal element in the Padé table preserve all quadratic first integrals of the original
Hamiltonian system.
τ −1 τ
Fτ = I − J −1 Hzz (z m+1 ) I + J −1 Hzz (z m )
2 2
is non-symplectic in general. By a nonlinear transformation [Dah59,QZZ95] ,
h h
ξ k = ρ(z k ) = z k + f (z k ), ξ k+1 = ρ(z k+1 ) = z k+1 + f (z k+1 ), (3.3)
2 2
and the trapezoidal scheme can be transformed into a symplectic Euler centered
scheme
h
ξ k + ξ k+1 = z k + z k+1 + f (z k ) + f (z k+1 ) .
2
Applying (3.2) to the above formula, we get
ξ k + ξ k+1
By taking z k+1 = in the second equation of (3.3), we obtain
2
% &
ξ k + ξ k+1 h ξ k + ξ k+1
ξ k+1 = + f ,
2 2 2
i.e., % &
ξ k + ξ k+1
ξ k+1 = ξ k + hf ,
2
which is a Euler centered scheme.
Theorem 3.3. The trapezoidal scheme (3.2) preserves the following symplectic
structure[WT03] :
h2
J + Hzz (z)JHzz (z), (3.4)
4
i.e.,
% k+1 & % &
∂z h2 k+1 k+1 ∂ z k+1 h2
k
J + H zz (z )JHzz (z ) k
= J+ Hzz (z k )JHzz (z k ).
∂z 4 ∂z 4
Proof. The proof can be easily obtained by direct calculation using nonlinear trans-
form of (3.3) to (1.7).
Remark 3.4. For the canonical system with general separable Hamiltonian, H(p, q) =
U (p) + V (q), and we have
dq dp
= −Vq (q), = Up (p), (3.5)
dt dt
1 m+1 1
(p − pm ) = −Vq (q m+ 2 ),
τ
(3.6)
1 m+1+ 1 1
(q 2 − q m+ 2 ) = U (pm+1 ).
p
τ
4.4 Explicit Symplectic Scheme for Hamiltonian System 203
! !
pm pm+1
The transition Fτ : 1 → 1 has the Jacobian:
q m+ 2 q m+1+ 2
5 6−1 5 6
I O I −τ L
Fτ = .
−τ M I O I
1
From Proposition 1.7, it is symplectic, but with M = Upp (pm+1 ), L = Vqq (q m+ 2 ).
3
Property 3.5. Let f (p, q) = p Bq be a conservation law of (3.5). Then, (pk+1 ) Bq k+ 2
1
= (pk ) Bq k+ 2 is a conservation law of the difference scheme (3.6) also.
Proof. Indeed, because f (p, q) is a conservation law of the original Hamiltonian sys-
tem
Bp, Up (p) = 0, Bq, Vq (q) = 0,
we get
% &
3
q k+ 2 − q k+ 2
1
, Bpk+1 = Up (pk+1 ), Bpk+1 = 0,
τ
% &
pk+1 − pk 1 1 1
, Bq k+ 2 = Vq (q k+ 2 ), Bq k+ 2 = 0.
τ
The oldest and simplest difference scheme is the explicit Euler method. Usually, it is
not symplectic for general Hamiltonian systems. It is interesting to ask: under what
condition of Hamiltonian systems, can the explicit Euler method become symplectic?
In fact, the explicit Euler scheme should be the phase flow of a system (i.e., exact
solution) to be symplectic. Most of the important Hamiltonian systems can be decom-
posed into the sum of these simple systems. Then, the composition of the Euler method
acting on these systems yields a symplectic method, which is also explicit. These sys-
tems are called symplectically separable. So classical separable Hamiltonian systems
are symplectically separable. In this section, we will prove that any polynomial Hamil-
tonian is symplectically separable.
204 4. Symplectic Difference Schemes for Hamiltonian Systems
z = EH
τ
z := z + τ JHz (z), (4.1)
τ
where EH = 1 + τ JHz . Usually, the scheme (4.1) is non-symplectic. However, it is
symplectic for a specific kind of Hamiltonian system, called a system with nilpotent
of degree 2.
[FW98]
Definition 4.1. A Hamiltonian system is nilpotent of degree 2 if it satisfies
Evidently, H(p, q) = φ(p) or H(p, q) = ψ(q), which represents inertial flow and
stagnant flow, are nilpotent of degree 2 since for H(p, q) = φ(p),
5 65 65 6 5 65 6
φpp O O −I φp φpp O O
Hzz (z)JHz (z) = = = O,
O O I O O O O φp
iff
CJC T = O. (4.4)
4.4 Explicit Symplectic Scheme for Hamiltonian System 205
Proof. Since
JHzz (z)JHz (z) = JC T φuu (CJC T φu (Cz)), (4.5)
the sufficient condition is trivial.
We now prove the necessity. If
JC T CJC T Cz = O, ∀ z,
i.e.,
JC T CJC T C = O.
Left multiplying by C and right multiplying JC T by this equation, we get:
(CJC T )3 = O.
Lemma 4.4. Let C = (A, B); then CJC T = O, if and only if AB T = BAT .
z = EH
τ
z = Eφτ z = z + τ JHz (z) = z + τ JC T φu (Cz)
m
H(z) = Hi (z), Hi (z) = φi (Ci z) = φ(Ai p + Bi q), (4.6)
i=1
206 4. Symplectic Difference Schemes for Hamiltonian Systems
with k-fold rotational symmetry in a phase plane[2,4] are not separable in the conven-
if k = 1, 2, 4. Otherwise
tional sense they are symplectically separable, since every
2π i 2π i
term cos p cos + q sin is nilpotent of degree 2 according to Theorem 4.3.
k k
For example, for k = 3,
2π 2π 4π 4π
H3 (p, q) = cos p + cos p cos + q sin + cos p cos + q sin
3 3 3 3
% √ & % √ &
1 3 1 3
= cos p + cos p− q + cos − p − q ,
2 2 2 2
4.4 Explicit Symplectic Scheme for Hamiltonian System 207
Define
P1 (x, y) : = (x + y)n − xn − y n
= C1n xn−1 y 1 + C2n xn−2 y 2 + · · · + C2n x2 y n−2 + C1n x1 y n−1 ,
which is separable, and the right side consists of mixed terms; P1 is a linear combina-
tion of 3 terms (x + y)n , xn , and y n .
P1 (x, 2y) = 2C1n xn−1 y 1 + 22 C2n xn−2 y 2 + · · · + 2n−2 C2n x2 y n−2 + 2n−1 C1n x1 y n−1 ,
2P1 (x, 1y) = 2C1n xn−1 y 1 + 2C2n xn−2 y 2 + · · · + 2C2n x2 y n−2 + 2C1n x1 y n−1 .
Define
P2 (x, y) : = P1 (x, 2y) − 2P1 (x, y)
= (22 − 2)C2n xn−2 y 2 + · · · + (2n−2 − 2)C2n x2 y n−2
+(2n−1 − 2)C1n x1 y n−1 ,
1 1
(x + y)2m+1 + (x − y) − x2m+1
2 2
= C22m+1 x2m−1 y 2 + C42m+1 x2m−3 y 4 + · · · + C2m
2m+1 xy
2m
,
1 1
(x + y)2m+1 − (x − y)2m+1 − y 2m+1
2 2
= C12m+1 x2m y + C32m+1 x2m−2 y 3 + · · · + C2m−1 2 2m−1
2m+1 x y ,
1 1
(x + y)2m + (x − y)2m − x2m − y 2m
2 2
= C22m x2m−2 y 2 + C42m x2m−4 y 4 + · · · + C2m−2
2m x2 y 2m−2 ,
1 1
(x + y)2m − (x − y)2m
2 2
= C12m x2m−1 y + C32m x2m−3 y 3 + · · · + C2m−1
2m xy 2m−1 ,
1 1 1 1 1
xy = (x + y)2 − x2 − y 2 = (x + y)2 − (x − y)2 .
2 2 2 4 4
4.5 Energy-conservative Schemes by Hamiltonian Difference 209
m
gree m = mi and with degree mi in variable ui . A and B are diagonal matrices
i=1
of order n:
⎛ ⎞ ⎛ ⎞
a1 0 ··· 0 b1 0 ··· 0
⎜ 0 a2 ··· 0 ⎟ ⎜ 0 b2 ··· 0 ⎟
⎜ ⎟ ⎜ ⎟
A=⎜ .. .. .. ⎟, B=⎜ .. .. .. ⎟,
⎝ . . . ⎠ ⎝ . . . ⎠
0 0 · · · an 0 0 · · · bn
However, these schemes are not symplectic. For simplicity, we illustrate the cases
only when n = 2. Let z = z m , z̄ = z m+1 .
1 1
(p̄1 − p1 ) = − {H(p1 , p2 , q̄1 , q2 ) − H(p1 , p2 , q1 , q2 )},
τ q̄1 − q1
1 1
(p̄2 − p2 ) = − {H(p̄1 , p2 , q̄1 , q̄2 ) − H(p̄1 , p2 , q̄1 , q2 )},
τ q̄2 − q2
1 1
(5.1)
(q̄1 − q1 ) = {H(p̄1 , p2 , q̄1 , q2 ) − H(p1 , p2 , q̄1 , q2 )},
τ p̄1 − p1
1 1
(q̄2 − q2 ) = {H(p̄1 , p̄2 , q̄1 , q̄2 ) − H(p̄1 , p2 , q̄1 , q̄2 )}.
τ p̄2 − p2
By addition and cancellation, we have energy conservation for the arbitrary Hamil-
tonian H(p̄1 , p̄2 , q̄1 , q̄2 ) = H(p1 , p2 , q1 , q2 ).
Since the proposed energy conservative schemes based on Hamiltonian differenc-
ing only have the first order accuracy, Qin[Qin87] first proposed another more symmet-
ric form in 1987, which possesses the second order accuracy. Independently, Itoh and
Abe[IA88] also proposed the same schemes in 1988.
For simplicity, we consider only the case n = 2, and the following difference
schemes are given:
d p1 1 H(p1 , p2 , q̄1 , q2 ) − H(p1 , p2 , q1 , q2 ) 1 H(p1 , p̄2 , q̄1 , q̄2 ) − H(p1 , p̄2 , q1 , q̄2 )
=− −
dt 4 Δq1 4 Δq1
1 H(p̄1 , p2 , q̄1 , q2 ) − H(p̄1 , p2 , q1 , q2 ) 1 H(p̄1 , p̄2 , q̄1 , q̄2 ) − H(p̄1 , p̄2 , q1 , q̄2 )
− − ,
4 Δq1 4 Δq1
d q1 1 H(p̄1 , p2 , q̄1 , q2 ) − H(p1 , p2 , q̄1 , q2 ) 1 H(p̄1 , p̄2 , q̄1 , q̄2 ) − H(p1 , p̄2 , q̄1 , q̄2 )
= +
dt 4 Δp1 4 Δp1
1 H(p̄1 , p2 , q1 , q2 ) − H(p1 , p2 , q1 , q2 ) 1 H(p̄1 , p̄2 , q1 , q̄2 ) − H(p1 , p̄2 , q1 , q̄2 )
+ + ,
4 Δ p1 4 Δp1
d p2 1 H(p̄1 , p2 , q̄1 , q̄2 ) − H(p̄1 , p2 , q̄1 , q2 ) 1 H(p1 , p2 , q1 , q̄2 ) − H(p1 , p2 , q1 , q2 )
=− −
dt 4 Δ q2 4 Δq2
1 H(p̄1 , p̄2 , q̄1 , q̄2 ) − H(p̄1 , p̄2 , q̄1 , q2 ) 1 H(p1 , p̄2 , q1 , q̄2 ) − H(p1 , p̄2 , q1 , q2 )
− − ,
4 Δq2 4 Δ q2
d q2 1 H(p̄1 , p̄2 , q̄1 , q̄2 ) − H(p̄1 , p2 , q̄1 , q̄2 ) 1 H(p1 , p̄2 , q1 , q̄2 ) − H(p1 , p2 , q1 , q̄2 )
= +
dt 4 Δp2 4 Δ p2
1 H(p̄1 , p̄2 , q̄1 , q2 ) − H(p̄1 , p2 , q̄1 , q2 ) 1 H(p1 , p̄2 , q1 , q2 ) − H(p1 , p2 , q1 , q2 )
+ + .
4 Δ p2 4 Δ p2
This chapter discusses the construction of the symplectic difference schemes via gen-
erating function and their conservation laws.
σα : M (m) −→ N (m),
|Cα M + Dα | = 0. (1.2)
!
Aα Bα
Proposition 1.2. Let α ∈ GL(2m), and the inverse α−1 = , then
Cα Dα
|Cα M + Dα | = 0 iff |M C α − Aα | = 0,
(1.3)
|Aα M + Bα | = 0 iff |B α − M Dα | = 0.
σα (M ) = (M C α − Aα )−1 (B α − M Dα ). (1.4)
i.e.,
214 5. The Generating Function Method
Aα Aα + Bα C α = Aα Aα + B α Cα = Im ,
Cα B α + Dα Dα = C α Bα + Dα Dα = Im ,
(1.5)
Aα B α + Bα Dα = Aα Bα + B α Dα = O,
Cα Aα + Dα C α = C α Aα + Dα Cα = O,
we obtain the following identities:
5 65 α 6 5 6
I −M A Bα Aα − M C α B α − M Dα
= ,
Cα Dα Cα Dα O I
5 65 6 5 6 (1.6)
I −M Aα Bα Aα − M C α B α − M Dα
= .
Aα Bα Cα Dα I O
In addition, we have:
5 6 5 65 6
I −M I O I −M
= ,
Cα Dα Cα I O Cα M + Dα
5 6 5 65 6 (1.7)
I −M I O I −M
= .
Aα Bα Aα I O Aα M + Bα
i.e.,
(B α − M Dα )(Cα M + Dα ) = (M C α − Aα )(Aα M + Bα ).
Expanding it and using the conditions (1.5), we know that it holds.
(C α N + Dα )(Cα M + Dα ) = I, (1.9)
hence
|C α N + Dα | = 0 iff |Cα M + Dα | = 0,
5.2 Symplectic, Gradient Mapping and Generating Function 215
Proof.
(C α N + Dα )(Cα M + Dα )
= (C α (Aα M + Bα )(Cα M + Dα )−1 + Dα )(Cα M + Dα )
= (C α Aα + Dα Cα )M + C α Bα + Dα Dα
=I (by (1.5)),
which is (1.9). The first equation of (1.10) can be obtained from (1.4) and the second
equation can be derived from (1.1).
Combining (1.2) and (1.3) together, we obtain the following four mutually equiv-
alent transversality conditions:
|Cα M + Dα | = 0, (1.11)
|M C α − Aα | = 0, (1.12)
|C α N + Dα | = 0, (1.13)
|N Cα − Aα | = 0, (1.14)
where
N = σα (M ) = (Aα M + Bα )(Cα M + Dα )−1 ,
M = σα−1 (N ) = (Aα N + B α )(C α N + Dα )−1 .
Moreover, the linear fractional transformation σα from {M ∈ M (m) | |Cα M +
Dα | = 0} to {N ∈ M (m) | |C α N + Dα | = 0} is 1-1 surjective.
n
n
"=
Ω d zi ∧ d zi+n − d zi ∧ d zi+n , (2.1)
i=1 i=1
216 5. The Generating Function Method
O
J2n
where the corresponding matrix is given by J"4n = . We denote
O −J2n
" 4n = (R4n , J"4n ).
R
On the other hand, R4n has its standard symplectic structure:
2n
Ω= d wi ∧ d w
i+2n , (2.2)
i=1
where (w1 , · · · , w2n , w 2n )T represents its coordinate. The corresponding ma-
1 , · · · , w
trix is given by 5 6
O I2n
J4n = .
−I2n O
We denote manifold R4n = (R4n , J4n ).
Now we first review some notations and facts of the symplectic algebra. Every
4n × 2n matrix of rank 2n can be represented as:
5 6
A1
A= ∈ M (4n, 2n), A1 , A2 ∈ M (2n),
A2
defines a 4n-dim subspace {A} spanned by its 2n column vectors. Evidently, {A} =
{B} iff ∃P ∈ GL(2n) such that
5 6 5 6
A1 P B1
AP = B, i.e., = .
A2 P B2
X
1
A 2n-dim subspace {X} = of R4n , X1 , X2 ∈ M (2n), is a J4n -Lagrangian,
X2
if
X T J4n X = O,
i.e.,
X1T X2 − X2T X1 = O X1T X2 ∈ Sm(2n).
or
X
1
According to Siegel[Sie43] , we call such a 4n × 2n matrix X = a symmetric
X2
X
1 N
pair. Moreover, if |X2 | = 0, then X1 X2−1 = N ∈ Sm(2n) and = .
X2 I
Y
1
Similarly, a 2n-dim subspace {Y } = is J˜4n -Lagrangian, if
Y2
Y T J˜4n Y = O,
i.e.,
Y1T J2n Y1 = Y2T J2n Y2 ,
5.2 Symplectic, Gradient Mapping and Generating Function 217
Y1
the 4n×2n matrix Y = is called a symplectic pair. |Y2 | = 0 implies Y1 Y2−1 =
Y2
Y1 M
M ∈ Sp(2n), and = .
Y2 I
A
Bα
Theorem 2.1. A transformation α = α
∈ GL(4n) carries every J˜4n -
Cα Dα
Lagrangian subspace into a J4n -Lagrangian subspace if and only if α ∈ CSp(J˜4n , J4n ),
i.e.,
αT J4n α = μJ˜4n , for some μ = μ(α) = 0. (2.3)
Proof. The “if” part is obvious, we need only to prove the “only if”
5 part. 6
J2n J2n
Taking α0 ∈ Sp(J˜4n , J4n ) (which always exists), e.g., α0 = 1 1 ,
I2n I2n
2 2
we have
CSp(J˜4n , J4n ) = CSp(4n) · α0 .
Therefore, it suffices to show that if α carries every J4n - Lagrangian subspace into
J4n -Lagrangian subspace, then α ∈ CSp(4n), i.e.,
Set P = AT T
α Dα − Cα Bα , then the above equation becomes
SP = P T S, ∀ S ∈ Sm(2n).
AT T
α Dα − Cα Bα = μI.
So
5 6T 5 65 6
T
Aα Bα O I Aα Bα
α J4n α =
Cα Dα −I O Cα Dα
5 6
AT T
α Cα − Cα Aα AT T
α Dα − Cα Bα
=
BαT Cα − DαT Aα BαT Dα − DαT Bα
5 6
O I
= μ = μJ4n ,
−I O
α ∈ GL(4n) implies μ = 0.
Aα Bα
The inverse matrix of α is denoted by α−1 = . By (2.3), we have
Cα Dα
AT T T T
α Cα − Cα Aα = μJ, Aα Dα − Cα Bα = O,
(2.4)
BαT Cα − DαT Aα = O, BαT Dα − DαT Bα = −μJ,
A Bα
Theorem 2.2. Let α = α
∈ CSp(J˜4n , J4n ). The linear fractional trans-
Cα Dα
formation σα :{M ∈ Sp(2n) | |Cα M +Dα | = 0}→ {N ∈ Sm(2n) | |C α N +Dα | =
0} is one to one and onto.
i.e.,
(M T AT T T T T T T
α + Bα )(Cα M + Dα ) = (M Cα + Dα )(Aα M + Bα ).
O = M T (AT T T T
α Cα − Cα Aα )M + Bα Dα − Dα Bα
+M T (AT T T T
α Dα − Cα Bα ) + (Bα Cα − Dα Aα )M
= M T JM − J,
By the way, we can get the original symplectic mapping, if f (w) or φ(w) is ob-
tained from Theorem 2.5.
5.3 Generating Functions for the Phase Flow 221
g 0 = identity, g t1 +t2 = g t1 ◦ g t2 ,
and if z0 is taken as an initial condition, then z(t) = g t (z0 ) is the solution of (3.1)
with the initial value z0 .
Theorem 3.1. Let α ∈ CSp(J˜4n , J4n ). Let z → z = g(z, t) be the phase flow of
the Hamiltonian system (3.1) and M0 ∈ Sp(2n). Set G(z, t) = g(M0 z, t) with Ja-
cobian M (z, t) = Gz (z, t). It is a time-dependent canonical map. If M0 satisfies the
transversality condition (1.2), i.e.,
|Cα M0 + Dα | = 0, (3.2)
then there exists, for sufficiently small |t| and in (some neighborhood of) R2n , a time-
dependent gradient map w → w = f (w, t) with Jacobian N (w, t) = fw (w, t) ∈
Sm(2n) satisfying the transversality condition (1.13) and a time-dependent generat-
ing function φα,H (w, t) = φ(w, t), such that
(3.4) is the most general Hamilton–Jacobi equation for the Hamiltonian system
(3.1) with the linear transformation α.
Proof. Since g(z, t) is differentiable with respect to z and t, so is G(z, t). Condition
(3.2) implies that for sufficiently small |t| and in some neighborhood of R2n ,
= −μH(Aα w(w,
t) + B α w). (3.8)
222 5. The Generating Function Method
∂ wi ∂wj
∂ wi2n
∂H
= − d wj ∧ d wi + − d t ∧ d wi . (3.9)
∂ wj ∂ wi ∂t ∂ wi
i<j i=1
∂w
Since N (w, t) = fw (w, t) = is symmetric, the first term of (3.9) is zero.
∂w
Notice that z = G(z, t) = g(M0 z, t),
d z d g(M0 z, t)
= = J −1 ∇ H(G(z, t)). (3.10)
dt dt
So G(z, t) is the solution of the following initial-value problem:
⎧
⎨ d z = J −1 ∇ H( z ),
dt
⎩ z(0) = M z.
0
= Aα G(z, t) + Bα z,
w w = Cα G(z, t) + Dα z,
it follows that
dw dw
= Aα J −1 ∇H(
z ), = Cα J −1 ∇H(
z ).
dt dt
dw ∂w dw
∂w
Since = + , combining these equations, we obtain
dt ∂ w dt ∂t
∂w
∂w
= Aα − Cα J −1 ∇H(
z ).
∂t ∂w
On the other hand,
T T
∂w
∇w H(w, t) = H w (w, t) = μ − Hz · Aα + Bα
∂w
α T T α T
∂w
= −μ (B ) + (A ) ∇H( z)
∂w
∂
w
= Aα J −1 − Cα J −1 ∇ H(
z ) by (2.5) and N ∈ Sm(2n)
∂w
∂w
= .
∂t
ω1 = w
d w + H d t = d φ(w, t),
i.e.,
f (w, t) = ∇w φ(w, t),
∂
φ(w, t) = −μH Aα ∇w φα,H (w, t) + B α w .
∂t
Therefore, the theorem is completed.
This is the generating function and H.J. equation of the first kind.
⎡ ⎤
O O −In O
⎢ O −In O O ⎥
(II) α=⎢ ⎣ O
⎥ , μ = 1, M0 = I, |Cα M0 + Dα | = 0;
O O In ⎦
In O O O
q
w= , φ = φ(q, p, t);
p
p φ
q
w=− = , φt = −H( p, −φp).
q φp
This is the generating function and H.J. equation of the second kind.
⎡ ⎤
−J2n J2n
(III) α=⎣ 1 1
⎦, μ = 1, M0 = I, |Cα M0 + Dα | = 0;
I2n I2n
2 2
1
w = (z + z), φ = φ(w, t);
2
1
= J(z − z) = ∇φ,
w φt = −H w − J −1 ∇φ .
2
Theorem 3.2. Let H(z) depend analytically on z. Then φα,H (w, t) is expressible as
a convergent power series in t for sufficiently small |t|, with recursively determined
coefficients:
224 5. The Generating Function Method
∞
φ(w, t) = φ(k) (w)tk , (3.11)
k=0
1 T
φ(0) (w) = w N0 w, N0 = (Aα M0 + Bα )(Cα M0 + Dα )−1 , (3.12)
2
φ(1) (w) = −μ(α)H(E0 w), (3.13)
E0 = Aα N0 + B α = M0 (Cα M0 + Dα )−1 .
If k ≥ 1,
2n
μ(α) 1
k
(k+1)
φ (w) = − Hzi1 ,···,zim (E0 w)
k+1 m!
m=1 i1 ,···,im =1 j1 +···+jm =k
jl ≥1
where Hzi1 ,···,zim (E0 w) is the m-th partial derivative of H(z) w.r.t. zi1 , · · · , zim ,
evaluated at z = E0 w and Aα ∇φ(jl ) (w) i is the il -th component of the column
l
vector Aα ∇φ(jl ) (w).
Proof. Under our assumption, the generating function φα,H (w, t) depends analyti-
cally on w and t in some neighborhood of R2n and for small |t|. Expand it as a power
series as follows:
∞
φ(w, t) = φ(k) (w)tk .
k=0
∞
∇φ(w, t) = ∇φ(k) (w)tk , (3.15)
k=0
∞
∂
φ(w, t) = (k + 1)tk φ(k+1) (w). (3.16)
∂t
k=0
By (3.15),
∇φ(0) (w) = ∇φ(w, 0) = f (w, 0) = N0 w.
1
So we can take φ(0) (w) = wT N0 w. We denote E0 = Aα N0 + B α . Then
2
∞
Aα ∇φ(w, t) + B α w = E0 w + Aα ∇φ(k) (w)tk .
k=1
Substitutes it in H Aα ∇φ(w, t) + B α w and expanding at z = E0 w, we get
5.3 Generating Functions for the Phase Flow 225
H(Aα ∇φ(w, t) + B α w)
* ∞
+
(k)
= H E0 w + A ∇φ (w)t
α k
k=1
∞ 2n ∞
1
= H(E0 w) + tj1 +···+jm Hzi1 ,···,zim
m=1
m ! i ,···,i =1 j ,···,j =1
1 m 1 m
∞ 2n
1
= H(E0 w) + tk Hzi1 ,···,zim
m ! i ,···,i =1
m=1 1 m k≥m j1 +···+jm =kjl ≥1
Substituting this formula into the R.H.S. of (3.4), and (3.5) into the L.H.S. of (3.4),
then comparing the coefficients of tk on both sides, we obtain the recursions Equations
(3.13) and (3.14).
In the next section when we use generating functions φα,H to construct difference
schemes we always assume M0 = I. For the sake of convenience, we restate Theorem
3.1 and Theorem 3.2 as follows.
Theorem 3.3. Let α ∈ CSp(J˜4n , J4n ). Let z → z = g(z, t) be the phase flow of the
Hamiltonian system (3.1) with Jacobian M (z, t) = gz (z, t). If
|Cα + Dα | = 0,
then there exists, for sufficiently small |t| and in (some neighborhood of) R2n , a time-
dependent gradient map w → w = f (w, t) with Jacobian N (w, t) = fw (w, t) ∈
Sm(2n) satisfying the transversality condition (1.13) and a time-dependent generat-
ing function φα,H (w, t) = φ(w, t) such that
Theorem 3.4. Let H(z) depend analytically on z. Then φα,H (w, t) is expressible as
a convergent power series in t for sufficiently small |t|, with the recursively determined
coefficients:
∞
φ(w, t) = φ(k) (w)tk ; (3.22)
k=0
1 T
φ(0) (w) = w N0 w, N0 = (Aα + Bα )(Cα + Dα )−1 ; (3.23)
2
φ(1) (w) = −μ(α)H(E0 w), E0 = (Cα + Dα )−1 . (3.24)
If k ≥ 1,
2n
μ(α) 1
k
(k+1)
φ (w) = − Hzi1 ,···,zim (E0 w)
k+1 m! m=1 i1 ,···,im =1 j1 +···+jm =k
jl ≥1
Theorem 4.1. Using Theorems 3.3 and 3.4, for sufficiently small τ > 0 as the time-
step, we define
m
ψ (m) (w, τ ) = φ(i) (w)τ i , m = 1, 2, · · · . (4.1)
i=0
(m)
Proof. Since ψ (m) (w, 0) = φ(w, 0), so ψww (w, 0) = φww (w, 0) = fw (w, 0) =
N (w, 0) satisfies the transversality condition (1.13), i.e., |C α N (w, 0) + Dα | = 0.
Thus for sufficiently small τ and in some neighborhood of R2n , N (m) (w, τ ) =
(m)
ψww (w, τ ) satisfies the transversality condition (1.13), i.e., |C α N (m) (w, τ ) + Dα | =
0. By Theorem 4.1, the gradient mapping w → w = f˜(w, τ ) = ∇ψ (m) (w, τ ) defines
implicitly a time-dependent canonical mapping z → z = g̃(z, τ ) by the equation
Aα z + Bα z = ∇ψ (m) (Cα z + Dα z, τ ).
Here we use the matrix notation instead of the component notation in Theorem 3.4.
Hzz denotes the Hessian matrix of H, and all derivatives of H are evaluated at z =
E0 w.
Type (II). Constructing symplectetic scheme by the second kind of the generating
function
228 5. The Generating Function Method
⎡ ⎤ ⎡ ⎤
O O −In O O O O In
⎢ O −In O O ⎥ ⎢ O −I O O ⎥
α=⎢ ⎣ O
⎥ , αT = α−1 = ⎢ n ⎥.
O O In ⎦ ⎣ −In O O O ⎦
In O O O O O In O
q p
w= , w=− ,
p q
O I O I O O
N0 = − , E0 = , Aα E0T = − ,
I O I O I O
φ(1) (w) = −H(
p, q),
1
n
φ(2) (w) = − (Hqi Hpi )(
p, q),
2
i=1
1 n
φ(3) (w) = − (Hpi pj Hqi Hqj + Hqi qj Hpi Hpj + Hqi pj Hpi Hqj ),
6
i,j=1
∂H
where H(z) = H(p1 , · · · , pn , q1 , · · · , qn ), Hzi = .
∂ zi
a. The first order scheme.
Evidently, (4.4) is an explicit difference scheme of 1-st order of accuracy. If we set q’s
1
at half-integer times t = k + τ , then (4.4) becomes
2
1
pk+1
i = pki − τ Vqi (q k+ 2 ),
k+ 12 +1 k+ 12
i = 1, · · · , n. (4.6)
qi = qi + τ Upi (pk+1 ),
τ 2
n
pk+1
i = pki − τ Hqi (pk+1 , q k ) − Hqj Hpj qi (pk+1 , q k )
2
j=1
τ3
n
− Hpl pj Hql Hqj + Hql qj Hpl Hpj + Hpl qj Hql Hpj qi
(pk+1 , q k ),
6
l,j=1
τ 2
n
qik+1 = qik + τ Hpi (pk+1 , q k ) + Hqj Hpj pi (pk+1 , q k )
2
j=1
τ3
n
+ Hpl pj Hql Hqj + Hql qj Hpl Hpj + Hpl qj Hql Hpj p (pk+1 , q k ),
6 i
l,j=1
where i = 1, · · · , n.
Type (III). Constructing symplectetic scheme by Poincaré type generating function
⎡ ⎤ ⎡ ⎤
1
−J2n J2n J2n I2n
α=⎣ 1 ⎦ , α−1 = ⎢ ⎣
2 ⎥
⎦. (4.7)
1 1
I2n I2n − J2n I2n
2 2 2
1
w = (
z + z), = J(z − z).
w (4.8)
2
N0 = 0, E0 = I, Aα E0T + E0 AαT = 0. (4.9)
φ(0) = φ(2) = φ(4) = 0, (4.10)
1
φ(1) (w) = −H ( z + z) , (4.11)
2
1
φ(3) (w) = (∇H)T JHzz J∇H, (4.12)
24
ψ (2) (w, τ ) = −τ H, (4.13)
τ3
ψ (4) (w, τ ) = −τ H + (∇H)T JHzz J∇H. (4.14)
24
230 5. The Generating Function Method
It is not difficult to show that the generating function φ(w, t) of type (III) is odd in
t. Hence, Theorem 4.1 leads to a family of canonical difference schemes of arbitrary
even order accuracy.
5 6
−J2n J2n
Theorem 4.2. Let α = 1 1 . For sufficiently small τ > 0 as the time-
I2n I2n
2 2
step, we define
m
ψ (2m) (w, τ ) = φ(2i−1) (w)τ 2i−1 , m = 1, 2, · · · . (4.17)
i=1
6
o(tm+1 )
?
−1 ∇ψ
g m (z, t) α f˜(w, t) ψ(w)
5.5 Further Remarks on Generating Function 231
z = EH,c
s
z: z = z + sJHz (c
z + (1 − c)z), (5.1)
1
with real number c being unconditionally symplectic if and only if c = , which
2
corresponds to the centered Euler scheme
z + z
z = z + sJHz . (5.2)
2
These simple propositions illustrate a general situation: apart from some very rare
exceptions, the vast majority of conventional schemes are non-symplectic. However, if
we allow c in (5.1) to be a real matrix of order 2n, we get a far-reaching generalization:
(5.1) is symplectic iff
1
c= (I2n + J2n B), B T = B, cT J + Jc = J. (5.3)
2
For H(p, q) = φ(p) + ψ(q), the above schemes P and Q reduce to explicit schemes.
A matrix α of order 4n is called a Darboux matrix if
αT J4n α = J˜4n ,
% & % &
O −I2n J2n O
J4n = , J˜4n = ,
I2n O O −J2n
% & % &
a b a1 b1
α= , α−1 = .
c d c1 d1
Every Darboux matrix induces a (linear) fractional transform between symplectic and
symmetric matrices
Sp(2n) −→ Sm(2n),
σα :
σα (S) = (aS + b)(cS + d)−1 = A for |cS + d| = 0
232 5. The Generating Function Method
Sm(2n) −→ Sp(2n),
σα−1 :
σα−1 (A) = (a1 A + b1 )(c1 A + d1 )−1 = S for |c1 A + d1 | = 0,
or alternatively
ag(z) + bz = (∇φ) cg(z) + dz ,
where φ is called the generating function of Darboux type α for the symplectic oper-
ator g.[FQ91] Then
or alternatively
a1 ∇φ(w) + b1 (w) = g(c1 ∇φ(w) + d1 w), (5.6)
where g is called the symplectic operator of Darboux type α for the generating func-
tion φ.
For the study of symplectic difference scheme, we may narrow down the class of
Darboux matrices to the subclass of normal Darboux matrices, i.e., those satisfying
a + b = 0, c + d = I2n . The normal Darboux matrices α can be characterized as
* + * +
a b J −J 1
α= = , c = (I + JB), B T = B, (5.7)
c d c I −c 2
* + * +
−1
a1 b1 (c − I)J I
α = = . (5.8)
c1 d1 cJ I
every Hamiltonian H with its phase flow etH and for every normal Darboux matrix α,
we get the generating function φ(w, t) = φtH (w) = φtH,α (w) of normal Darboux
type α for the phase flow of H by
∇φtH,α = (JetH − J) ◦ (cetH + I − c)−1 for small |t|. (5.9)
φtH,α satisfies the Hamilton–Jacobi equation
∂
φ(w, t) = −H(w + a1 ∇φ(w, t)) = −H(w + c1 ∇φ(w, t)) (5.10)
∂t
and can be expressed by Taylor series in |t|:
∞
φ(w, t) = φ(k) (w)tk , |t| small enough. (5.11)
k=1
dz
= J∇H(z). (6.1)
dt
Suppose
z = gH
s
(z) (6.2)
y = S −1 ◦ gH
s
◦ S(y). (6.4)
On the other hand, the algorithm g s can be applied to system (6.3) directly and the
corresponding scheme is
y = gH̃
s
(y). (6.5)
Naturally, one can ask if (6.4) and (6.5) are the same. This introduces the following
concept.
S −1 ◦ gH
s
◦ S = gH◦S
s
, ∀ S ∈ G;
S −1 gH
s
◦ S = gH◦S
s
, ∀ S ∈ Sp(2n).
In practice, the second case is more common. Generally speaking, numerical al-
gorithms depend on the coordinates, i.e., they are locally represented. But many nu-
merical algorithms may be independent of the linear coordinate transformations.
5.6 Conservation Laws 235
F ◦ gH
s
= F + c, c is a constant (6.6)
s
if and only if gH is etF -invariant.
s
Proof. We first assume that the symplectic algorithm gH is etF -invariant, i.e.,
e−t
F ◦ gH ◦ eF = gH◦et ,
s t s
∀ t ∈ R. (6.7)
F
Since F is a first integral of the Hamiltonian system (6.1) with the Hamiltonian
H, H is also the first integral of the Hamiltonian system (5.6) with the Hamiltonian
F , i.e.,
H ◦ etF = H. (6.8)
It follows from (5.6) and (6.8) that
e−t
F ◦ gH ◦ eF = gH ,
s t s
i.e.,
s −1
etF = (gH ) ◦ etF ◦ gH
s
. (6.9)
Differentiating (6.9) with respect to t at point 0 and noticing that
d etF
= J∇F,
dt t=0
we get
s −1
J∇F = (gH )∗ J∇F ◦ gH
s
. (6.10)
s
Since gH is symplectic, i.e.,
s −1 s T
(gH )∗ J = J(gH )∗ ,
we have
s T
J∇F = J(gH )∗ ∇F ◦ gH
s
= J∇(F ◦ gH
s
),
then
s T
∇F = (gH )∗ ∇F ◦ gH
s
= ∇(F ◦ gH
s
).
It follows that
F ◦ gH
s
= F + c. (6.11)
s
We now assume that F is conserved by gH , i.e., (6.6) is valid. Then noticing that
s −1 s −1
the phase flows of the vector fields J∇F and (gH )∗ J∇F ◦ gH s
are etF and (gH ) ◦
eF ◦ gH respectively, we can get (5.6) similarly, i.e., gH is eF -invariant.
t s s t
Symplectic invariant algorithms are invariant under the symplectic group Sp(2n)
and hence invariant under the phase flow of any quadratic Hamiltonian.
236 5. The Generating Function Method
Corollary 6.3. Symplectic invariant algorithms for Hamiltonian systems preserve all
quadratic first integrals of the original Hamiltonian systems up to a constant.
s
If a symplectic scheme has a fixed point, i.e., there is a point z such that gH (z) = z,
then the constant c = 0 and the first integral is conserved exactly. Since linear schemes
always have the fix point 0, we have the following result.
Corollary 6.4. Linear symplectic invariant algorithms for linear Hamiltonian sys-
tems preserve all quadratic first integrals of the original Hamiltonian systems.
Example 6.5. Centered Euler scheme and symplectic Runge–Kutta methods are sym-
plectic invariants. Hence they preserve all quadratic first integrals of system (6.1) up
to a constant.
Example 6.6. Explicit symplectic scheme (4.5), and other explicit symplectic schemes
(2.1) – (2.4) considered in Chapter 8 are invariant under the linear symplectic transfor-
mations of the form diag (A−T , A), A ∈ GL(n). Thus they preserve angular momen-
tum pT Bq of the original Hamiltonian systems, since their infinitesimal symplectic
matrices are diag (−B T , B), B ∈ gl(n).
F (
z ) = F (z), or F ◦ gH,α
s
= F. (6.13)
1 T 1
z − z) A(
JB( z − z)T (AJB − BJA)(
z − z) = ( z − z) = 0, ∀ z, z ∈ R2n .
2 4
Combining it with (6.14), we have
T
z + (I − c)z A(
c z − z) = 0.
Using (5.13), it becomes
∞
T
z + (I − c)z
c AJ tj ∇φ(j) c
z + (I − c)z = 0.
j=1
we have
m
wT A(
z − z) = − sj wT AJ∇φ(j) (w) = −AJ∇ψ(w) = 0,
j=1
since
1 T 1 T 1
wT A(
z − z) = z A
z− z − z)T (AJB − BJA)(
z Az + ( z − z)
2 2 2
1 1 T
= zT Az− z Az.
2 2
Therefore, the theorem is completed.
We list some of the most important normal Darboux matrices c, the type matrices
B, together with the corresponding form of symmetric matrices A of the conserved
1
quadratic invariants F (z) = z T Az:
2
1
c = I − c = I, B = O, A arbitrary,
2
I
n O O
−In
c= , B= ,
O O −In O O b
b arbitrary;
O O O I A = , angular
bT O
c= , B= n
, momemtum.
O In In O
I a b T
1n ±In a = a, bT = −b;
c= , B = ∓I2n , A= ,
∓In In
2 −b a Hermitian type.
I ±I I a
1 n O b aT = a,
c= , B=± ,A= , T
2 ±I O O −In −b −a b = −b.
238 5. The Generating Function Method
Apart from the first integrals of the original Hamiltonian systems, a linear sym-
plectic algorithm has its own quadratic first integrals. For the linear Hamiltonian sys-
tem
dz
= Lz, L = JA ∈ sp(2n) (6.15)
dt
1
with a quadratic Hamiltonian H(z) = z T Az, AT = A, let us denote its linear
2
symplectic algorithm by
z = gH
s
(z) = G(s, A)z, G ∈ Sp(2n). (6.16)
Let us assume that the scheme (6.16) is of order r. Then G(s) has the form
G(s) = I + sL(s),
s 2 s2 sr−1 r
L(s) = L + L + L3 + · · · + L + O(sr ).
2! 3! r!
So (6.16) becomes
"
z = esL(s) z.
This is the solution z(t) of the linear Hamiltonian system
dz " "
= L(s)z, L(s) ∈ sp(2n), (6.17)
dt
with the initial value z(0) = z 0 evaluated at time s. The symplectic numerical solution
"
z k = Gk (s)z 0 = eksL(s) z 0
is just the solution of system (6.17) at discrete points ks, k = 0, ±1, ±2, · · ·. Hence,
for sufficiently small s, scheme (6.16) corresponds to a perturbed linear Hamiltonian
system (6.17) with the Hamiltonian
1 T −1
" s) = 1 z, J −1 L(s)z
H(z, " = z J Lz + O(sr ) = H(z) + O(sr ). (6.18)
2 2
It is well-known that the linear Hamiltonian system has n functionally independent
quadratic first integrals. So does the scheme (6.15). The following
" i (z, s) = 1 z T J −1 L
H " 2i−1 (s)z, i = 1, 2, · · · , n (6.19)
2
are the first integrals of the perturbed system (6.17), therefore, of scheme (6.16), which
approximate the first integrals of system (6.15)
1
Hi (z) = z T J −1 L2i−1 z, i = 1, 2, · · · , n
2
5.7 Convergence of Symplectic Difference Schemes 239
i (z, s) = z T J −1 Gi (s)z,
H i = 1, 2, · · · , n.
In this section, we shall prove that all symplectic schemes for Hamiltonian systems
constructed by generating functions are convergent, if τ → 0.
A normal Darboux matrix, which will be introduced in the next chapter, has the
form
5 6 ⎡ ⎤
Aα Bα J −J
α= =⎣ 1 1
⎦ , B T = B,
Cα Dα (I + JB) (I − JB)
2 2
⎡ ⎤
5 6 1 (7.2)
A α
B α (JBJ − J) I
−1 ⎢ 2 ⎥
α = =⎣ ⎦,
C α Dα 1
(JBJ + J) I
2
i.e.,
1 1
= J z − Jz,
w w = (I + JB) z + (I − JB) z, B T = B. (7.3)
2 2
Let z → z = g(z, t) be the phase flow of the Hamiltonian systems (5.7); it is a time
dependent canonical map. There exist, for sufficiently small |t| and in (some neigh-
borhood of) R2n , a time-dependent gradient map w → w = f (w, t) with Jacobian
fw (w, t) ∈ Sm(2n) (i.e.,everywhere symmetric) and a time-dependent generating
function φ = φα,H , such that
On the other hand, for a given time-dependent scalar function ψ(w, t) : R2n × R →
R, we can obtain a time-dependent canonical map g"(z, t). If ψ(w, t) approximates the
generating function φα,H (w, t) of the Hamiltonian system (5.7), then g"(z, t) approxi-
mates the phase flow g(z, t). For sufficiently small τ > 0 as the time step, define
m
φ(m) = φ(k) (w)τ k , (7.5)
k=1
1
where φ(1) (w) = −H(w), and for k ≥ 0, Aα = (JBJ − J),
2
2n
−1 1
k
φ(k+1) (w) = Hzi1 · · · zim (w)
k+1 m!m=1 i1 ,···,in =1
· Aα ∇ φ(j1 ) (w) i1
· · · Aα ∇ φ(jm) (w) i . (7.6)
m
j1 +···+jm =k
Then, ψ (m) (w, τ ) is the m-th approximation of φα,H (w, τ ) , and the gradient map,
we have z k = g"τk .
First, we prove that the convergence holds locally. We begin by showing that for
any z0 , the iterations are defined for z"t/k
n
(n ≤ k), if t is sufficiently small. Indeed,
in the neighborhood of z0 , g"τ (z) = z + o(τ ), thus, if g"lt (z) (l = 1, 2, · · · , n − 1) is
k
defined for z in the neighborhood of z0 ,
5.7 Convergence of Symplectic Difference Schemes 241
g"nt (z) − z = g"nt (z) − g"n−1
t (z) + g"n−1
t (z) − g"n−2
t (z) + · · · + g" kt (z) − z
k k k k k
t t
= o + ··· + o = o(t).
0 k 12 k
3
n
≤ z1 − z2 + C F (s) d s,
0
using Gronwall inequality, we have
F (t) = g(z1 , t) − g(z2 , t) ≤ eC|t| z1 − zn ,
gt (z) − g"kt = g kt (z) − g"kt
k k k
k−1
= gt g kt (z) − g k−1
t g" kt (z) + g k−2
t g kt (y1 )
k k k
−g k−2
t g" kt (y1 ) + · · · + g k−1
t g kt (yl−1 )
k k
−g t k−1
g" kt (yl−1 ) + · · · + g kt (yk−1 ) − g" kt (yk−1 ),
k
general theory of generating function roughly reads as follows. Let a normal Darboux
matrix be
5 6 ⎡ ⎤
Aα Bα J −J
α= =⎣ 1 1
⎦ , B T = B,
Cα Dα (I + JB) (I − JB)
2 2
⎡ 1 ⎤
5 6 (JBJ − J) I
Aα Bα ⎢ 2 ⎥
α−1 = =⎣ ⎦.
α α 1
C D (JBJ + J) I
2
i.e.,
1 1
= J z − Jz,
w z + (I − JB)z,
w = (I + JB) B T = B. (8.7)
2 2
Let z → z = g(z, τ ) be the phase flow of the Hamiltonian system (8.2). It is a
time-dependent canonical map. There exists, for a sufficiently small τ and in (some
neighborhood of) R2n+2 , a time-dependent gradient mapping w → w = f (w, τ ) with
Jacobian fw (w, t) ∈ Sm(2n + 2) (i.e., symmetric everywhere) and a time-dependent
generating function φ = φα,K (w, τ ), such that
∂φ
= −K Aα ∇φ(w) + B α w ,
∂τ
f (w, τ ) = ∇φα,K (w, τ ), (8.8)
Aα g(z, t) + Bα z ≡ (∇φ) Cα g(z, t) + Da z, t .
On the other hand, for a given time-dependent scalar function ψ(w, t) : R2n+2 ×R →
R, we can get a time-dependent canonical map g"(z, τ ). If ψ(w, τ ) approximates the
generating function φα,K (w, τ ) of the Hamiltonian system (8.2), g"(z, t) approximates
the phase flow g(z, t).
For a sufficiently small s > 0 as the time-step, define
m
ψ (m) = φ(k) (w) sk , (8.9)
k=1
where
1
φ(1) (w) = −K(w), Aα = (JBJ − J).
2
For k ≥ 0,
244 5. The Generating Function Method
2n
−1 1
k
(k+1)
φ (w) = Kzi1 ···zim (w)
k+1 m!
m=1 i1 ,···,im =1
· Aα ∇φ(j1 ) (w) i1
· · · Aα ∇φ(jm ) (w) im .
j1 + · · · + jm = k
jl = 0
(8.10)
Then ψ (m) (w, s) is the m-th approximation of φα,K (w, s), and the gradient mapping
i.e.,
pk+1 + pk q k+1 + q k tk+1 + tk
pk+1
i = pki − sHqi , ,
2 2 2
s3
− (Hpj pl qi Hqj Hql + 2Hpj pl Hqj qi Hql − 2Hqj pl qi Hpj Hql
24
−2Hqj pl Hpj qi Hql − 2Hqj pl Hpj Hql qi + 2Hqj ql Hpl qi Hpj + Hqj ql qi Hpl Hpj
−2Hqj qi Hpj t − 2Hqj Hpj qi t + 2Hpj qi Hqj t + 2Hpj Hqj qi t + Hqi tt ),
k+1
p + pk q k+1 + q k tk+1 + tk
qik+1 = qik + sHpj , ,
2 2 2
s3
+ (Hpj pl pi Hqj Hql + 2Hpj pl Hqj pi Hql − 2Hqj pl pi Hpj Hql
24
−2Hqj pl Hpj pi Hql − 2Hqj pl Hpj Hql pi + 2Hqj pl Hpl qi Hpj + Hqj ql pi Hpl Hpj
−2Hqj pi Hpj t − 2Hqj Hpj pi t + 2Hpj pi Hqj t + 2Hpj Hqj pi t + Hpi tt ),
k+1
p + pk q k+1 + q k tk+1 + tk
hk+1 = hk − sHt , ,
2 2 2
3
s
− (Hpj pl t Hqj Hql + 2Hpj pl Hqj t Hql − 2Hqj pl t Hpj Hql
24
−2Hqj pl Hpj t Hql − 2Hqj pl Hpj Hql t + 2Hqj ql Hpl t Hql + Hqj ql t Hpl Hpj
−2Hqj t Hpj t − 2Hqj Hpj tt + 2Hpj t Hqj t + 2Hpj Hqi t + Httt ),
tk+1 = tk + s.
(8.16)
Let ! !
O I p
B=− , w= .
I O q
We have
φ(1) = −K(w),
1
φ(2) = − (Kqi Kpi )(w),
2
1
φ(3) = − (Kpj pl Kqj Kql + Kqj ql Kpj Kpl + Kqj pl Kpj Kql )(w), or
6
1 (8.17)
φ(2) = − (Hqi Hpj + Ht )(w),
2
1
φ(3) = − (Hpj pl Hqi Hql + Hql Hqj pl Hpj + Hql Hpl t
6
+Hqj ql Hpj Hpl + 2Hpj Hqj t + Htt ).
pk+1
i = pki − sHqi (pk+1 , q k , tk ),
tk+1 = tk + s.
In the previous chapter, we constructed the symplectic schemes of arbitrary order via
generating function. However the construction of generating functions is dependent
on the chosen coordinates. One would like to know under what circumstance will the
construction of generating functions be independent of the coordinates. The generating
functions are deeply associated with the conservation laws, so it is important to study
their properties and computations.
α : T ∗ Rn × T ∗ Rn −→ T ∗ (Rn × Rn )
maps the symplectic structure (1.2) to the standard one (1.3). In particular, α maps La-
grangian submanifolds in (R4n , J"4n ) to Lagrangian submanifolds Lg in (R4n , J4n ).
Suppose that α satisfies the transversality condition of g Chapter 5, Equation (1.2),
then 5 6 7
dφg (ω)
∗ 2n
Lg = , ω∈T R , (1.5)
ω
φg is called generating function of g. We call this generating map α (linear case) or
α∗ (nonlinear case) Darboux transformation, in other words, we have the following
definition.
Definition 1.1. A linear map
5 6
Aα Bα
α= , (1.6)
Cα Dα
Denote
Eα = Cα + Dα , Fα = Aα + Bα , (1.8)
then, we have:
φα,g : R2n −→ R,
such that
Aα g(z) + Bα z = ∇φα,g (Cα g(z) + Dα z),
i.e.,
(Aα g + Bα )(Cα g + Dα )−1 z = ∇φα,g (z).
dimM . Denote M ≡ {α ∈ M | Eα = In , Fα = 0} ⊂ M ∗ ⊂ M .
The following theorem answers the question on how to normalize a given Darboux
transformation.
252 6. The Calculus of Generating Functions and Formal Energy
such that β2 β1 α ∈ M .
Proof. We need only to take P = −Fα Eα−1 = −(Aα +Bα )(Cα +Dα )−1 , T = Eα−1 ,
then
5 T 65 65 6
Eα O I −Fα Eα−1 Aα Bα
β2 β1 α =
O Eα−1 O I Cα Dα
!
Aβ2 β1 α Bβ2 β1 α
=
Cβ2 β1 α Dβ2 β1 α
5 6
Eα (Aα − Fα Eα−1 Cα ) Eα (Bα − Fα Eα−1 Dα )
= α1 = . (2.1)
Eα−1 Cα Eα−1 Dα
It’s easy to verify that
!
Aβ O
Because α1 ∈ M , we have Dβ = I2n , Bβ = O, i.e., β = . Since
! Cβ I2n
I2n O
β ∈ Sp(4n), we have β = , Q ∈ Sm(2n). Thus:
Q I2n
5 6⎡ J −J2n
⎤
I2n O 2n
α1 = ⎣ ⎦
1 1
Q I2n I2n I2n
2 2
⎡ ⎤
J2n −J2n
=⎣ 1 1
⎦
I2n + QJ2n I2n − QJ2n
2 2
⎡ ⎤
J2n −J2n
=⎣ 1 1
⎦,
(I2n + V ) (I2n − V )
2 2
From the following theorem, we can show that the normalization condition is nat-
ural.
We first prove the “ only if ” part of the theorem. When taking τ = 0, we have
∂Gτ (z)
A = −Hz ((C + D)z).
∂τ τ =0
Since
∂Gτ (z)
= J −1 Hz (z),
∂τ τ =0
we have
AJ −1 Hz (z) = −Hz ((C + D)z), ∀ H, z.
Take special form H(z) = z T b, and substitute it into above equation, we have
AJ −1 b = −b, ∀ b,
z = (C + D)z, ∀ z =⇒ C + D = I.
A + B = O, A = −J, C + D = I, (2.3)
then
A(Gτ (z) − z) = −τ Hz (CGτ (z) + Dz),
A = −J, τ = 0 =⇒ Gτ (z)τ =0 = z.
Proof. Since
⎡ ⎤ 5 6
T T
(T −1 ) Aα (T −1 ) Bα AβT α BβT α
βT α = ⎣ ⎦= ,
T Cα T Dα CβT α DβT α
we have
Aα g(z) + Bα z = ∇φα,g ◦ (Cα g(z) + Dα z), (3.2)
and
T T
(T −1 ) Aα g(z) + (T −1 ) Bα z = ∇φβT α,g ◦ (T Cα g(z) + T Dα z)
⇐⇒ Aα g(z) + Bα z = T (∇φβT α,g ) ◦ T (Cα g(z) + Dα z)
= ∇(φβT α,g ◦ T )(Cα g(z) + Dα z). (3.3)
Thus we obtain
φα,g ∼
= φβT α,g ◦ T
or
φα,g ◦ T −1 ∼
= φβT α,g
The theorem is proved.
!
S O "
Theorem 3.2. ∀ S ∈ Sp(2n), define γS = ∈ Sp(4n). Then we have
O S
φαγS ,g ∼
= φα,S◦g◦S −1 (3.4)
256 6. The Calculus of Generating Functions and Formal Energy
Proof. Since:
5 6 5 6
Aα S Bα S Aα γS Bα γS
αγS = = ,
Cα S Dα S Cα γS Dα γS
we have
Finally, we obtain:
∇φαγS,g = ∇φα,S◦g◦S −1 ,
i.e.,
φαγS,g ∼
= φα,S◦g◦S −1 .
The proof can be obtained.
!
I2n P
Theorem 3.3. Take β = ∈ Sp(4n), P ∈ Sm(2n), α ∈ Sp(J"4n , J4n ),
O I2n
then:
φβα,g ∼
= φα,g + ψp , (3.7)
1
where ψp = w P w ( function independent of g).
2
Proof. Since:
5 6 5 6 5 6
I2n P Aα Bα Aα + P Cα Bα + P Dα
βα = = ,
O I2n Cα Dα Cα Dα
Eβα = Eα , Fβα = Fα + P Eα ,
obviously,
we obtain
φβα,g − ψP ∼
= φα,g .
Analogically, we have:
!
I2n O
Theorem 3.4. If we take β = ∈ Sp(4n), Q ∈ Sm(2n), then
Q I2n
1
φα,g + (∇w φα,g (w)) Q(∇w φα,g (w)) ∼
= φβα,g (w + Q∇φα,g (w)). (3.12)
2
Theorem 3.5. We have the following relation:
φ⎡ ⎤ ∼
= −φ⎡ ⎤ . (3.13)
⎣
A B ⎦ ,g −1 ⎣
−B −A ⎦ ,g
C D D C
Proof. Since
Theorem 3.6. If
φ⎡ ⎤ ∼
= −φ⎡ ⎤ , ∀ g,
⎣
A B ⎦ ,g −1 ⎣
A B ⎦ ,g −1
C D C D
then
A + B = O, C = D.
258 6. The Calculus of Generating Functions and Formal Energy
Proof. Since
Ag(S −1 z) + Bz = ∇φα,g◦S −1 Cg(S −1 z) + Dz , ∀ z,
φ⎡ ⎤ ∼
= φ⎡ ⎤ .
⎣
A BS ⎦ ,g ⎣
A B ⎦ ,g◦S −1
C DS C D
Therefore, the theorem is completed.
Theorem 3.11. If
!
λI2n O
β= ∈ CSp(4n), α ∈ Sp(J"4n , J4n ), λ = 0,
O I2n
!
λA λB
βα= ∈ CSp(J"4n , J4n ), μ(βα) = λ,
C D
then we have
φ⎡ ⎤ ∼
= λφ⎡ ⎤ . (3.17)
⎣
λA λB ⎦ ,g ⎣
A B ⎦ ,g
C D C D
Proof. Since
then we have
260 6. The Calculus of Generating Functions and Formal Energy
φ⎡ ⎤ ∼
= λφ⎡ ⎤ .
⎣
λA λB ⎦ ,g ⎣
A B ⎦ ,g
C D C D
then we have:
φ⎡ ⎤ ∼
= λφ⎡ ⎤ ◦ λ−1 I2n . (3.18)
⎣
A B ⎦ ,g ⎣
A B ⎦ ,g
λC λD C D
Proof. Since
!
A B
∈ CSp(J"4n , J4n ),
λC λD
Ag(z) + Bz = ∇φ⎡ ⎤ (λCg(z) + λDz),
⎣
A B ⎦ ,g
λC λD
L.H.S = ∇φ⎡ ⎤ (Cg(z) + Dz),
⎣
A B ⎦ ,g
C D
⎛ ⎞
⎜ ⎟
R.H.S = ⎜
⎝φ
⎡ ⎤ ⎟ ◦ λI2n (Cg(z) + Dz)
⎠
⎣
A B ⎦ ,g
λC λD
⎛ ⎞
⎜ ⎟
= λ−1 ∇ ⎜
⎝φ
⎡ ⎤ ◦ λI2n ⎟
⎠ (Cg(z) + Dz),
⎣
A B ⎦ ,g
λC λD
hence
φ⎡ ⎤ ∼
= λ−1 φ⎡ ⎤ ◦ λI2n .
⎣
A B ⎦ ,g ⎣
A B ⎦ ,g
C D λC λD
Therefore, the theorem is completed.
6.4 Invariance of Generating Functions and Commutativity of Generator Maps 261
Before finishing this section, we will give two conclusive theorems which can
include the contents of the seven theorems given before. They are easy to prove and
the proofs are omitted here.
Let !
a b
α ∈ CSp(J"4n , J4n ), β ∈ CSp(J4n ), β = ,
c d
obviously
β α ∈ CSp(J"4n , J4n ), μ(βα) = λ(β)μ(α),
and then the following theorem.
Theorem 3.13. For φβα,g , we have
φβα,g (c∇w φα,g (w) + dw) (3.19)
'
∼ 1
= λ(β)φα,g (w) + w (d b)w + (∇w φα,g (w))
2
/
1
·(c b)w (∇w φα,g (wλ)) (c a)(∇w φα,g (w)) . (3.20)
2
If
(Aα M + Bα )(Cα M + Dα )−1 = (Aα M + Bα )(Cα M + Dα )−1 ,
∀ M ∼ I2n , M ∈ Sp(2n),
then
α = ±α.
262 6. The Calculus of Generating Functions and Formal Energy
Proof. Let
N0 = (Aα I + Bα )(Cα I + Dα )−1
= (Aα I + Bα )(Cα I + Dα )−1 .
Aβ − Dβ − N0 Cβ − Cβ N0 = εCβ = 0 =⇒ Cβ = 0,
thus Aβ = Dβ . !
◦ Aβ Bβ
From 1 , we have B = , Bβ = Bβ . Therefore
O Aβ
Aβ N A−1 −1
β = N − Bβ Aβ .
Aβ (N − N0 ) = (N − N0 )Aβ .
φα,g ∼
= φα,g , ∀ g ∈ Sp-diff, g ∼ I2n =⇒ α = ±α.
φα,g ∼
= φα,g =⇒ Hessian(φα,g ) = (φα,g )ww
= (Aα g(z) + Bα )(Cα g(z) + Dα )−1 ,
6.4 Invariance of Generating Functions and Commutativity of Generator Maps 263
For the invariance of generating function φg (S) under S, one would like to expect
φα,S −1 ◦g◦S = φα,g ◦ S, ∀g ∼ I.
This is not true in general case. We shall study under what condition this is true for the
normalized Darboux transformation αV . The following theorem answers this ques-
tion.
Theorem 4.3. Let
⎡ ⎤
J2n −J2n
α = αV = ⎣ 1 1
⎦, ∀ V ∈ sp(2n), αV ∈ M ,
(I + V ) (I − V )
2 2
5 6
(S −1 )T O
S ∈ Sp(2n), βS = ∈ Sp(J4n ),
O S
5 6
S O
γS = ∈ Sp(J"4n ).
O S
F ◦ gH
t
= F ⇐⇒ {F, H} = 0 ⇐⇒ H ◦ gFt = H ⇐⇒ gH
t
= gF−S ◦ gH
t
◦ gFS .
Theorem 4.5. Let F be a conservation law of Hamiltonian system, then phase flow
t
gH (or symplectic schemes φτH ) keeps phase flow gFt with F (or φτF ) invariant iff
F ◦ gH = F + C. Let F ∈ F, g ∈ Sp-diff, then
F ◦ g = F + c =⇒ ∇F = ∇F ◦ g =⇒ gFt = gFt ◦g
= g −1 ◦ gFt ◦ g
On the other hand, take the derivative of both sides of the following equation w.r.t. t
at t = 0,
gFt (z) = g −1 gFt (g(z)),
and notice that g∗ (z) ∈ Sp, g∗−1 J −1 = J −1 g∗T , we get
then we have
∇F = ∇F ◦ g =⇒ F ◦ g = F + c.
Therefore, the theorem is completed.
with the following property: if hs (z) converges, the phase flow ght s is a canoni-
cal transformation with Hamiltonian function hs (z), which is considered as a time-
independent Hamiltonian with s as a parameter and satisfies “equivalence condition”
ght s t=s = F s . (5.1)
Therefore hs (z) = hs (F s z), ∀z ∈ R2n , thus hs (z) is invariant under F s (for those
s, z in the domain of convergence of hs (z)).
The generating function with F s , the new Hamiltonian function and α, the Dar-
boux transformation is
∞
φF s ,α (w) : ψ(s, w) = sk ψ (k) (w). (5.2)
k=1
Assuming it converges, we associate the phase flow with the generating function
∞
hs (z) −→ ψht s ,α (w) : χ(t, s, w) = tk χ(k) (s, w),
k=1
(1)
χ (s, w) = −h(s, w). (5.3)
For k > 1,
k 2n
1
χ(k+1) (s, w) = − hwl1 ,···,wlm (s, w)
(k + 1)m!
m=1 l1 ,···,lm =1 k1 +···+km =k
·(A1 χ(k1)
w (s, w))l1 · · · (A1 χ(k
w
m)
(s, w))lm . (5.4)
∞
∞
∞
Let χ(k) (s, w) = si χ(k,i) (w), then χ(t, s, w) = tk si χ(k,i) (w). Then
i=0 k>1 i=0
∞
∞
k
1
si χ(k+1,i) (w) = si
(k + 1)m!
i=0 i=0 m=1
i0 + i1 + · · · + im = i
k1 + · · · + km = k
2n
(1,i )
· χwl1 0,···,wlm (w)(A1 χ(k
w
1 ,i1 )
(w))l1 · · ·
l1 ,···,lm =1
·(A1 χ(k
w
m ,im )
(w))lm . (5.5)
266 6. The Calculus of Generating Functions and Formal Energy
Thus
k 2n
1 (1,i )
χ(k+1,i) (w) = χwl1 0,···,wlm (w)
(k + 1)m!
m=1 i0 +i1 +···+im =i l1 ,···,lm =1
k1 +···+km =k
·(A1 χ(k
w
1 ,i1 )
(w))l1 · · · (A1 χ(k
w
m ,im )
(w))lm . (5.6)
Let
∞
∞
χ(1) (s, w) = si χ(1,i) (w) = −h(s, w) = − si hi (w),
i=0 i=0
(k+1,i)
so the coefficient χ can be obtained by recursion,
i.e.,
∞
∞
∞
sk si χ(k,i) (w) = sk ψ (k) (w),
k>1 i=0 k>1
∞
∞
si χ(k−j,j) (w) = si ψ (k) (w),
i=0 j=0 i=1
k−1
χ(k−j,j) (w) = ψ (k) , k = 2, 3, · · · , (5.9)
j=0
χ(1,0) = ψ (1) ,
k
χ(k+1−i,i) = ψ (k+1) , k = 2, 3, · · · .
i=0
So
6.5 Formal Energy for Hamiltonian Algorithm 267
ψ (k+1) χ(k+1,0)
and
χ(1,0) = ψ (1) ,
h(0) = −ψ (1) ,
So h(0) , h(1) , h(2) , · · · can be recursively determined by ψ (1) , ψ (2) , · · ·. So we get the
∞
formal power series hs = si h(i) (z), and in case of convergence, it satisfies
i=0
ght s t=s = F s .
We now give a special example to show how to calculate the formal energy. Let us
consider normal Darboux transformation with
268 6. The Calculus of Generating Functions and Formal Energy
5 6 5 6
O −I A1 B1
V = −E = , αV−1 = ,
−I O C1 D1
where 5 6
1 O O
A1 = (JV J − J) = .
2 −I O
Suppose we just use the first term of the generating function of the generating map
αV , i.e., we just consider the first order scheme
∞
F s ∼ ψ(s, w) = −sH(w) = sk ψ (k) .
i=1
Let us assume ψ (1) = −H(w). If ψ (2) = ψ (3) = · · · = 0, then χ(1,0) = ψ (1) = −H.
We need to calculate χ(2,0) . Since
5 6
(1,0)
Hp
χz =− ,
Hq
5 65 6 5 6
(1,0)
O O −Hp O
A1 χz = = ,
−I O −Hq Hp
5 6
(1,0)
Hpp Hpq
χzz = − .
Hqp Hqq
1 (1,0) (1,0)
= (χz ) A1 χz
2
5 6 5 6
1 Hp 0 1
=− = − Hq Hp .
2 Hq Hp 2
1
χ(2,0) + χ(1,1) = ψ (2) = 0 =⇒ χ(1,1) = −χ(2,0) = H Hp .
2 q
In order to obtain χ(1,2) , we first determine χ(3,0) and χ(2,1) , and for the latter we
need to caculate
6.5 Formal Energy for Hamiltonian Algorithm 269
⎡ ⎤
∂
n
⎢ ∂p Hqj Hpj ⎥
⎢ ⎥
(1,1) 1⎢ j=1 ⎥
χz = ⎢ ⎥
2⎢ ⎥
⎣ ∂
n
⎦
Hqj Hpj
∂q
j=1
5 6
1 Hpq Hp + Hpp Hq
= ,
2 Hqq Hp + Hqp Hq
5 6
(1,1)
O O (1,1)
A1 χz = χz
−I O
5 6
1 O
=− ,
2 Hpq Hp + Hpp Hq
(2,0) (1,1)
A1 χz = −A1 χz
5 6
1 O
= .
2 Hpq Hp + Hpp Hq
For k = 2, i = 0, we have
2n
2n
2n
1 1
χ(3,0) = χ(1,0)
zl (A1 χ(k
z
1 ,0)
)l1
3 1! 1
i0 +i1 =0 k1 =2 l1 =1
2n
+ χ(1,0) (k1 ,0)
zl ,zl (A1 χz )l1 (A1 χ(k
z
2 ,0)
)l2
1 2
l1 ,l2 =1 i0 +i1 +i2 =0 k1 +k2 =2
For k = 1, i = 1, we have
2n
2n
2n
1
χ(2,1) = χz(1,i
l
0)
(A1 χ(k
z
1 ,i1 )
)l1
2 1
i0 +i1 =1 k1 =1 l1 =1
1 (1,0) T (1,1) (1,1) T (1,0)
= χz A1 χz + χz A1 χz
2
1 1
= (H T Hpp Hq + HpT Hqq Hp ) + HqT Hpq Hp .
4 q 2
270 6. The Calculus of Generating Functions and Formal Energy
For k = 1, i = 2,
1 1
χ(1,2) = − − (HqT Hpp Hq + HpT Hqq Hp ) − HqT Hpq Hp
6 6
1 1
+ (HqT Hpp Hq + HpT Hqq Hp ) + HqT Hpq Hp
4 2
1
= − (H T Hpp Hq + HpT Hqq Hp + 4HqT Hpq Hp ).
12 q
t
Now, let H(z) be a time-independent Hamiltonian, let its phase flow be gH , and
let its generating function be
∞
φgH
t (w) = φ(t, w) = tk φ(k) (w).
k=1
Then we have
φ(1) (w) = −H(w),
for k ≥ 1,
k 2n
1 (1)
φ(k+1) (w) = φwl1 ···wlm (w)
(k + 1)m!
m=1 l1 ,···,lm =1 k1 +···+km =k
· A1 φw (w) l · · · A1 φ(k
(k1 )
1
w
m)
(w) lm . (5.13)
then
6.5 Formal Energy for Hamiltonian Algorithm 271
Using equation H = ψ (1) = φ(1) = χ(1,0) and (5.9), (5.14) ,we get
Applying equation
then
χ(2,2) = χ(3,2) = χ(4,2) = · · · = χ(k,2) = · · · = 0.
Finally
then
χ(2,m−1) = χ(3,m−1) = χ(4,m−1) = · · · = χ(k,m−1) = · · · = 0.
so we finally get
∞
h(s, z) = si χ(1,i) = H(z) + sm (ψ (m+1) − φ(m+1) ) + O (|s|m+1 ),
i=0
i.e.,
h(s, z) − H(z) = sm (ψ (m+1) (z) − φ(m+1) (z)) + O(|s|m+1 ).
ψ (m+1) = φ(m+1) = 0,
then
h(s, z) = H(z) − O(|s|m+1 ).
H ◦ φτ = H,
Theorem 6.1 (G–M theorem). [Ge88] There exists a function τ = τ (c, t)defined on a
neighborhoold of 0 ∈ R, such that
φτ (c,t) H=c = g t H=c .
This means that if we can find a symplectic scheme preserving energy, we can solve the
original Hamiltonian system equivalently by a reparametrization of time parameter in
t
phase flow gH . In general this is impossible.
The proof of above Theorem 6.1 bases on the following Lemma 6.2.
t t
Lemma 6.2. Let gA 1
, gA 2
be solutions of following systems of ODE respectively.
dx dx
= A1 (x, t), = c(t)A1 (x, t),
dt dt
where c(t) is function of t, then
t τ (t)
gA 1
= gA2 ,
Proof. omit.
which is tangent to the energy surface H = c. Its restriction to the level surface H = c
is
J −1 F1 (c, τ )Hz .
H=c
−1
It differs from the restriction of vector field J Hz to the level surface H = c only
by a constant F1 (c, τ ). By Lemma 6.2 the proof is completed.
All symplectic transformations that keep H invariant compose a group S(H).
S(H) is a rigidity under which S0 (H) is a contained connected support set of unit
transformation of group S(H). S0 (H) induces the level surface H = c by the set of
all transformation, denoted by S0 (H)|H=0 . Then S0 (H) is a curve
t
S0 (H) = {gH H=c
, t ∈ R}.
Note that the rigidity of S0 (H) exactly counteracts the existence of energy-preserving
symplectc scheme.
Bibliography
[Fen98] K. Feng: The calculus of generating functions and the formal energy for Hamiltonian
systems. J. Comput. Math., 16:481–498, (1998).
[FQ87] K. Feng and M.Z. Qin: The symplectic methods for the computation of Hamiltonian
equations. In Y. L. Zhu and B. Y. Guo, editors, Numerical Methods for Partial Differential
Equations, Lecture Notes in Mathematics 1297, pages 1–37. Springer, Berlin, (1987).
[FQ91a] K. Feng and M.Z. Qin: Hamiltonian Algorithms for Hamiltonian Dynamical Systems.
Progr. Natur. Sci., 1(2):105–116, (1991).
[FQ91b] K. Feng and M.Z. Qin: Hamiltonian algorithms for Hamiltonian systems and a com-
parative numerical study. Comput. Phys. Comm., 65:173–187, (1991).
[FQ03] K. Feng and M. Z. Qin: Symplectic Algorithms for Hamiltonian Systems. Zhejiang
Press for Science and Technology, Hangzhou, in Chinese, First edition, (2003).
[Ge88] Z. Ge: Symplectic geometry and its application in numerical analysis. PhD thesis,
Computer Center, CAS, (1988).
[Ge91] Z. Ge: Equivariant symplectic difference schemes and generating functions. Physica
D, 49:376–386, (1991).
[GM88] Z. Ge and J. E. Marsden: Lie–Poisson Hamilton–Jacobi theory and Lie–Poisson inte-
grators. Physics Letters A, pages 134–139, (1988).
[GW95] Z. Ge and D.L. Wang: On the invariance of generating functions for symplectic trans-
formations. Diff. Geom. Appl., 5:59–69, (1995).
Chapter 7.
Symplectic Runge–Kutta Methods
dz
= J −1 Hz , (1.2)
dt
or
dz
= J −1 Hz = f (z). (1.3)
dt
The s-stage R–K method for (1.3) has the following form:
s
z k+1 = z k + h bi f (Yi ),
i=1
s (1.4)
Yi = z k + h aij f (Yj ), 1 ≤ i ≤ s,
j=1
s
where ci = aij (i = 1, 2, · · · , s). Thus, a s-stage R–K method is determined
j=1
completely by the mathmatical tableau (1.6). Therefore this kind of expression is often
called the Butcher tableau (or form).
7.1 Multistage Symplectic Runge–Kutta Method 279
Definition 1.1. A symplectic R–K method is a R–K method whose transitional trans-
∂ z k+1
formation of (1.4), i.e., Jacobian matrix is symplectic.
∂z k
s
1
B(p) : bi ck−1
i = , k = 1(1)p,
k
i=1
s
cki
C(η) : aij ck−1
j = , k = 1(1)η,
k
j=1
s
bj (1 − ckj )
D(ζ) : bi ck−1
i aij = , j = 1(1)s, k = 1(1)ζ,
k
i=1
where A is s×s matrix, b and c are s×1 vectors of weights and abscissae, respectively.
Theorem 1.3. If the coefficients A, b, c of a R–K method satisfy B(p), C(η), D(ζ)
(p ≤ η + ζ + 1, and p ≤ 2η + 2), then the R–K method is of order p[HNW93] .
R–K method is based on high order quadrature rule. Thus one can derive a R–K
method of order s for any set of distinct abscissas ci (i = 1, · · · , s). A high order can
be obtained for the following special sets of abscissas:
1◦ Using shifted zeros of the Gauss–Legendre polynomial to obtain ci and con-
dition C(s) of Definition 1.2 to obtain the Gauss–Legendre method.
2◦ Using zeros of Radau polynomial:
ds−1 s
(1) x (x − 1)s−1 (left Radau),
dxs−1
ds−1 s−1
(2) x (x − 1)s (right Radau),
dxs−1
with condition D(s) of Definition 1.2 to obtain Radau I A method, or with condition
C(s) to obtain Radau II A method.
3◦ Using zeros of Lobatto polynomial
ds−2 s−1
s−2
x (x − 1)s−1
dx
1 1
0 −
4 4
0 1
2 1 5
1 3 4 12
1 3
4 4
Radau II A:
1 5 1
−
3 12 12
1 1
3 1
1
4 4
1
3 1
4 4
Lobatto III A:
0 0 0 0
0 0 0
1 5 1 1
−
1 1 2 24 3 24
1
2 2 1 2 1
1
1 1 6 3 6
2 2 1 2 1
6 3 6
Lobatto III B:
1 1
0 − 0
1 6 6
0 0
2 1 1 1
0
1 2 6 3
1 0
2 1 5
1 0
1 1 6 6
2 2 1 2 1
6 3 6
Lobatto III C:
7.1 Multistage Symplectic Runge–Kutta Method 281
1 1 1
0 −
1 1 6 3 6
0 −
2 2 1 1 5 1
−
1 1 2 6 12 12
1
2 2 1 2 1
1
1 1 6 3 6
2 2 1 2 1
6 3 6
We present a table of these conditions for methods which are based on high order
quadrature rule, see Table 1.1.
Table 1.1. The simplified conditions for s-stage method based on high order quadrature rule
Proof. Here we give our own proof[QZ92a] . To prove the scheme (1.4) is symplectic
when M = 0, we only need to verify the Jacobian matrix is symplectic. From the
scheme (1.4) we have
∂z k+1 s
∂Y
= I + h bi Df (Yi ) ki , (1.7)
∂z k ∂z
i=1
∂Yi s
∂Y
=I +h aij Df (Yj ) kj , 1 ≤ i ≤ s, (1.8)
∂z k ∂z
j=1
∂Yi
Denote Di = D f (Yi ), = Xi (i = 1, 2, · · · , s), and let f = J −1 Hz , then
∂z k
JDi + Di J = 0, (1.9)
and
% & * s + * s +
∂z k+1 ∂z k+1
J = J +h bi Di Xi J + hJ bi Di Xi
∂z k ∂z k
i=1 i=1
* + * +
s
s
+ h bi Di Xi J h bi Di Xi
i=1 i=1
s
= J +h bi [(Di Xi ) J + JDi Xi ]
i=1
* s + * s +
2
+h bi Di Xi J bi Di Xi .
i=1 i=1
s
(Xi ) JDi Xi = JDi Xi + h aij (Dj Xj ) JDi Xi .
j=1
s
+h bi [Xi Di JXi + Xi JDi Xi ]
i=1
s
s
−h bi h aij (Di Xi ) JDj Xj
i=1 j=1
s !
+h aij (Dj Xj ) JDi Xi
j=1
s
s
= J + h2 (bi bj − bi aij − bj aji )(Di Xi ) JDj Xj .
i=1 j=1
% &
∂ z k+1 ∂ z k+1
J = J,
∂ zk ∂ zk
∂z k+1
i.e., the Jacobian matrix of transitional mapping is symplectic.
∂z k
Remark 1.5. If R–K method is non-reducible, then condition M = 0 is also neces-
sary.
s
Let this method satisfy simplified conditions B(s) and C(s). Solve equations bi ck−1
i
i=1
1 1
s
= (1 ≤ k ≤ s) for bi (i = 1, · · · , s), and solve equations aij ck−1
j = cki (1 ≤
k kj=1
k ≤ s, 1 ≤ i ≤ s) for aij (i, j = 1, · · · , s). Then the scheme determined by bi and aij
is the only R–K method that has achieved 2s-order of accuracy. We listed Butcher’s
tableau for s ≤ 2 as follows
s = 1:
1 1
2 2
(1.12)
1
s = 2:
√ √
3− 3 1 3−2 3
6 4 12
√ √
3+ 3 3+2 3 1 (1.13)
6 12 4
1 1
2 2
It is easy to see that s = 1 is exactly the case of the Euler centered scheme:
% &
k+1 k 1 k k+1
z = z + hf (z + z ) . (1.14)
2
284 7. Symplectic Runge–Kutta Methods
It is not difficult to verify that both schemes (1.12) and (1.13) satisfy the conditions
M = 0, and hence are symplectic. Furthermore, we have the following conclusions:
Theorem 1.6. An s-stage Gauss–Legendre method is a symplectic scheme with 2s-
order of accuracy.
Proof. Since the scheme satisfies conditions D(s), C(s), B(2s), i.e.,
s
1 1 1
bi aij cl−1
i = bj (1 − clj ) = bj − bj clj
l l l
i=1
s
s
= bi bj cl−1
i − bj aji cl−1
i ,
i=1 i=1
which results in
s
(bi aij + bj aji − bi bj )cl−1
i = 0, l, j = 1, 2, · · · , s.
i=1
i
bi
where ci = bj−1 + (i = 1, · · · , s, b0 = 0).
j=1
2
Proof. Since the scheme is diagonally implicit, aij = 0 (j > i); to satisfy M = 0,
we have bi bj − bi aij − bj aji = 0 (i, j = 1, 2, · · · , s), which results in
bi
aij = bj , aii = , i = 1, · · · , s, i > j.
2
The theorem is proved.
Corollary 1.8. Explicit R–K method with any order does not satisfy condition M = 0.
Remark 1.9. Tableau (1.16) Cooper[Coo87] has discussed the condition (1.15) and con-
structed a method of family (1.16) with s = 3 and order 3.
s=2:
1 1
0
4 4
3 1 1 (1.18)
4 2 4
1 1
2 2
s=3:
1 1
a a
2 2
3 1
a a a
2 2
(1.19)
1 1
+a a a −a
2 2
a a 1 − 2a
Corollary 1.10. If s = 3, and the elements in Butcher tableau are taken in symmetri-
cal version (a11 = a33 ).
1 1
a a
2 2
1 1
a −a
2 2
(1.20)
1 1
1− a a 1 − 2a a
2 2
a 1 − 2a a
b1 b1
2 2
b2 b2
b1 + b1
2 2
b3 b3 (1.21)
b 1 + b2 + b1 b2
2 2
b4 b4
b 1 + b 2 + b3 + b1 b2 b3
2 2
b1 b2 b3 b4
s
1
bi c3i = , (1.26)
4
i=1
s
1
bi ci aij cj = , (1.27)
8
i,j=1
s
1
bi aij c2j = , (1.28)
12
i,j=1
s
1
bi aij ajk ck = . (1.29)
24
i,j,k=1
Now we have 8 equations with 4 unknowns. Luckily we find a set of solutions using
computer, which is
b1 = −2.70309412, b2 = −0.53652708,
b3 = 2.37893931, b4 = 1.8606818856.
Perhaps we can reduce the equations to the form of 4 equations with 4 unknowns and
s
get the exact solution. For an example, using bi = 1, bi bj − bi aij = 0 (i, j =
i=1
s
s
1
1, 2, · · · , s), we have bi ai,j = bi ci = . So we can remove Equation
2
i=1,j=1 i=1
(1.23) from the system. In an implementation of this R–K method, we rewrite it in the
following form:
b1 h
Y1 = z k + f (Y1 ),
2
b h
Y2 = 2Y1 − z k + 2 f (Y2 ),
2
b h (1.30)
Y3 = 2Y2 − (2Y1 − z k ) + 3 f (Y3 ),
2
b h
Y4 = 2Y3 − (2Y2 − 2Y1 + z k ) + 4 f (Y4 ),
2
z k+1 = 2Y4 − (2Y3 − 2Y2 + 2Y1 − z k ).
Corollary 1.11. This scheme (1.30) can be obtained by applying the implicit midpoint
scheme over 4 steps of length b1 h, b2 h, b3 h, b4 h. It has 4-th order accuracy.
Let
288 7. Symplectic Runge–Kutta Methods
1
1 z 4 + z0
z 4 = z 0 + b1 hf ,
2
2 1
2 1 z4 +z4
z 4 = z 4 + b2 hf ,
2
3 2 (1.31)
3 2 z4 +z4
z 4 = z 4 + b3 hf ,
2
3
3 z1 + z 4
z 1 = z 4 + b4 hf ,
2
Rewrite it in the following form:
1 % 1 &
z 4 + z0 b z 4 + z0
= z 0 + 1 hf ,
2 2 2
2 1 % 2 1 &
z4 +z4 1 b z4 +z4
= z 4 + 2 hf ,
2 2 2
3 2 % 3 2 & (1.32)
z4 +z4 2 b z4 +z4
= z 4 + 3 hf ,
2 2 2
3 % 3 &
z1 + z 4 3 b z1 + z 4
= z 4 + 4 hf .
2 2 2
Let
1 1 2
z0 + z 4 z4 +z4
= Y1 , = Y2 ,
2 2
2 3 3
z4 +z4 z 4 + z1
= Y3 , = Y4 ,
2 2
Exercise 1.12. Does there exist 5-stage diagonally implicit R–K method with 5th-
order accuracy?
The second order derivative can be expressed via Jacobian matrix. However, the
third-order derivative y (3) , can no longer be expressed via matrix and vector sym-
bol, not to mention the higher order derivative. This has motivated people to study
the structure of the Taylor expansion of high order derivatives and search for a better
symbol to simplify the Taylor expansion of high order derivatives. Then the rooted tree
theory[But87,Lam91,HNW93,SSC94] (With the tree roots skill to express high order deriva-
tive) emerged. Take y (3) as an example:
1 (3)
y = 1 f11 (1 f )2 + 1 f12 (1 f )(2 f ) + 1 f1 1 f1 (1 f ) + 1 f2 (2 f )
+1 f21 (2 f )(1 f ) + 1 f22 (2 f )2 + 1 f2 2 f1 (1 f ) + 2 f2 (2 f ) ,
2 (3)
y = 2 f11 (1 f )2 + 2 f12 (1 f )(2 f ) + 2 f1 1 f1 (1 f ) + 1 f2 (2 f )
+2 f21 (2 f )(1 f ) + 2 f22 (2 f )2 + 2 f2 2 f1 (1 f ) + 2 f2 (2 f ) . (1.36)
f (M ) (z)(K1 , K2 , · · · , KM )
m m m m
= ··· i
fj1 j2 ···jM j1 K1j2 K2 · · · jM
KM · ei , (1.37)
i=1 j1 =1 j2 =1 jM =1
and
1 , 0, · · · , 0]T ∈ Rm .
ei = [0, 0, · · · , 0, 0123
i
290 7. Symplectic Runge–Kutta Methods
i ∂ 3 (i f ) i ∂ 3 (i f )
f111 = ; f112 = i f121 = i f211 = ,
∂(1 z)3 ∂(1 z)2 ∂(2 z)
i = 1, 2.
i ∂ 3 (i f ) ∂ 3 (i f )
f122 = i f212 = i f221 = , i
f222 = ,
∂(1 z)∂(2 z)2 ∂(2 z)3
(3) The argument z simply denotes the vector with respect to whose component
we are performing the partial differentiations.
(4) An M times Frechet derivatives has M operands. This is an important prop-
erty to note.
Take m = 2, we have
Case M = 1,
2
2
f (1) (z)(K1 ) = i
fj1 (j1 K1 )ei
i=1 j1 =1
5 1 6
f1 (1 K1 ) + 1 f2 (2 K1 )
= , (1.39)
2
f1 (1 K1 ) + 2 f2 (2 K1 )
where
i ∂(i f ) i ∂(i f )
f1 = , f2 = , i = 1, 2.
∂(1 z) ∂(2 z)
Replace z with y, and K1 with f , (1.39) becomes
5 1 1 6
f1 ( f ) + 1 f2 (2 f )
f (1) (y)(f (y)) = = y (2) . (1.40)
2
f1 (1 f ) + 2 f2 (2 f )
Case M = 2,
2
2
2
f (2) (z)(K1 , K2 ) = i
fj1 j2 (j1 K1 )(j2 K2 )ei
i=1 j1 =1 j2 =1
5 1
6
f11 ( K1 )( K2 ) + 1 f12 (1 K1 )(2 K2 ) + 1 f21 (2 K1 )(1 K2 ) + 1 f22 (2 K1 )(2 K2 )
1 1
= .
2
f11 (1 K1 )(1 K2 ) + 2 f12 (1 K1 )(2 K2 ) + 2 f21 (2 K1 )(1 K2 ) + 2 f22 (2 K1 )(2 K2 )
5 1 6
f11 (1 f )2 + 2(1 f12 )(1 f )(2 f ) + 1 f22 (2 f )2
f (2) (y)(f (y), f (y)) = , (1.42)
2
f11 (1 f )2 + 2(2 f12 )(1 f )(2 f ) + 2 f22 (2 f )2
which is only part of the right side of (1.36), but not all. The absent terms are
5 1 1 1 6
f1 f1 ( f ) + 1 f2 (2 f ) + 1 f2 2 f1 (1 f ) + 2 f2 (2 f )
2
. (1.43)
f1 1 f1 (1 f ) + 1 f2 (2 f ) + 2 f2 2 f1 (1 f ) + 2 f2 (2 f )
Now if we replace the operand f (y) with f (1) (y)(f (y)) in (1.40), the result is exactly
(1.43). Hence, shortening the notation as in (1.41), (1.36) can be written as
Thus we have seen that y (2) is a single Frecht derivative of order 1, and that y (3) is
a linear combination of Frecht derivatives of order 1 and 2. In general, y (p) turns out
to be a linear combination of Frecht derivatives of order up to p − 1. The components
in such linear combination are called elementary differentials.
Definition 1.14. The elementary differential Fs : Rm → Rm of f , and their order
are defined respectively by
1◦ f is only elementary differential of order 1, and
2◦ if Fs (s = 1, 2, · · · , M ) are elementary differential of order rs , then the
Frecht derivative
F (M ) (F1 , F2 , · · · , FM ), (1.45)
is an elementary of order
M
1+ rs . (1.46)
s=1
3◦ The order of the elementary differential (1.47) is, by (1.46), the sum of the
M
orders of the elementary differentials plus 1, i.e., 1 + rs , where 1 is “for the brack-
s=1
ets”.
Order 1 has only one elementary differential, i.e., f .
Order 2 has only one elementary differential, i.e., f (1) (f ) = {f }.
Order 3 has two elementary differentials, i.e.,
2. Labeled graph
Let n be a positive integer. A labeled n-graph g is a pair {V, E} formed by a set V
collection with card (V ) = n, and a set E of unordered pairs (v, w) as a collection of
elements, of which v, w are point of the set V , and v = w. Therefore g may be empty.
V and E elements are known as the vertices and edges respectively. Two vertices v, w
is said to be the adjacent if (v, w) ∈ E. Fig. 1.1 shows labeled graph for n = 2, 3, 4.
◦ ◦ ◦ ◦ ◦
i j i j k
n=2 n=3
l k
◦ ◦ j k
◦ ◦
◦ j ◦ ◦ ◦ ◦
i j k l
◦ ◦
i l
◦ i n=4
A graph can have many different types of label. However for the same graph, there
exists a isomorphic mapping χ between two different labels. For an example, for one
of the graphs in Fig. 1.1, depicted also in Fig. 1.2, we can take two types of label, as
shown in Fig. 1.3.
◦ ◦ ◦ ◦
i j k l
◦ ◦ ◦ ◦ ◦ ◦ ◦ ◦
i j k l m n p q
We take mapping χ to be
χ : i −→ m, j −→ n, k −→ p, l −→ q,
i.e.,
χ : V1 −→ V2 , V1 = {i, j, k, l}, V2 = {m, n, p, q},
then
χ: (i, j) −→ (m, n),
(j, k) −→ (n, p),
(k, l) −→ (p, q),
i.e.,
χ: E1 −→ E2 ,
E1 = {(i, j), (j, k), (k, l)},
E2 = {(m, n), (n, p), (p, q)}.
Therefore χ : Lg1 → Lg2 , where Lg1 = {V1 , E1 }, Lg2 = {V2 , E2 } are two different
labels in Fig. 1.3. In fact, if there exists a isomorphic mapping between two labels
g1 , g2 , they can be regarded as two types of labels for the same tree. Therefore a
graph is an equivalent class, which consists of a variety of different labeled graphs
corresponding to different types of label. These labeled graphs are equivalent, i.e.,
there exists an isomorphic mapping between them.
3. Relationship between rooted tree and elementary differential
Next we can see that there is a 1 to 1 correspondence between elementary and trees.
(1) Let f be the unique elementary differential of order 1. Then f corresponds
to the unique tree of order 1, which consists of a single vertex.
(2) If the elementary differential Fs of order rs (s = 1, 2, · · · , M ) corre-
sponds to trees ts of order rs (s = 1, 2, · · · , M ), then the elementary differential
M
M
{F1 , F2 , · · · , FM } of order 1 + rs corresponds to the tree of order 1 + rs ,
1 s=1
obtained by grafting the M trees Fs (s = 1, 2, · · · , M ) onto a new root.
◦ ◦
◦ ◦ ◦
Example 1.16. If F1 ∼ t1 , F2 ∼ t2 = , F3 ∼ t3 = ◦ , then
◦ ◦
◦◦ ◦ ◦
◦ ◦ ◦ ◦
{F1 , F2 , F3 } ∼ ◦ ◦ .
◦
We need a notation to represent trees similar to the notation for elementary differ-
ential. All trees can be labeled with combination of the symbol of τ for the unique tree
of order 1 (consisting of a single node) and the symbol [· · ·], meaning we have grafted
the trees appearing between the brackets onto a new root. We shall denote n copies of
[[[···[ ]···]]]
tree t1 by tn1 , ktimes by [k , and ktimes by ]k . For example
294 7. Symplectic Runge–Kutta Methods
◦
t1 = [τ ]= ,
◦
◦
@ A @ A @ ◦ ◦
t2 = τ [τ ] = [τ ]τ = τ [τ ]2 = ,
◦
@ A
t3 = [t1 , t22 ] = [τ ] [τ [τ ]], ] [τ [τ ] ]
◦
◦ ◦
◦ ◦
◦ ◦
= [2 τ ] [τ [τ ]2 [τ [τ ]3 =
◦ ◦ .
◦
◦
Definition 1.17. The order r(t), symmetry σ(t) and density (tree factorial) γ(t) are
defined by
Let α(t) (tree multiplicity) be the number of essentially different ways of labeling
the vertices of the tree t with the integers 1, 2, · · · , r(t) such that labels are monotone
increase. Essentially different labeling is illustrated in the following examples:
◦ ◦ 2 3◦ 4 4 2◦ 3
◦ ◦ ◦ and ◦ ◦,
Example 1.18. t1 = [τ 3 ] = , its labeling trees are
◦ ◦ ◦
1 1
are not regarded as essentially different labelings, hence α(t1 ) = 1.
4 4
◦ ◦ ◦
@ A ◦ ◦ 2◦ ◦3 3◦ ◦2
Example 1.19. t2 = τ [τ ] = , its labeling trees are ,
◦ ◦ ◦
1 1
3
◦
4◦ ◦2
and are regarded as essentially different labelings, and α(t2 ) = 3.
◦
1
From above, we have a easy way of computing α(t), namely
r(t) !
α(t) = . (1.48)
σ(t)γ(t)
7.1 Multistage Symplectic Runge–Kutta Method 295
F (τ ) = f,
(1.49)
F ([t1 , t2 , · · · , tM ]) = {F (t1 ), F (t2 ), · · · , F (tM )}.
The proof of the following two theorems was established by Butcher in 1987 [But87] .
y (2) = {f },
y (3) = {f 2 } + {2 f }2 ,
y (4) = {f 3 } + 3{f {f }2 + {2 f 2 }2 + {3 f }3 .
Let us define the right side of Equation (1.4) to be yn (h), which is then expanded as a
Taylor series about h = 0,
1
y(xn+1 ) = y(xn ) + hy (1) (xn ) + h2 y (2) (xn ) + · · · , (1.51)
2
where
dq
q
y n (h) = α(t)γ(t)φ(t)F (t). (1.52)
dh h=0
r(t)=q
s
φi (τ ) = aij ,
j=1
s
φi ([t1 , t2 , · · · , tM ]) = aij φj (t1 )φj (t2 ) · · · φj (tM ).
j=1
◦
[τ ] {f } = f f 2 1 2 1 aij cj bi ci
◦ j i
◦ ◦
[τ 2 ] {f 2 } = f (f, f ) 3 2 3 1 aij c2j bi c2i
◦ j i
◦
◦ [[τ ]] {2 f }2 = f f f 3 1 6 1 aij ajk ck bi aij cj
◦ j k i,j
◦
◦ ◦ [τ 3 ] {f 3 } = f (f, f, f ) 4 6 4 1 aij c3j bi c3i
◦ j i
◦
◦ ◦
[τ [τ ]] {f {f }2 = f (f, f f ) 4 1 8 3 aij cj ajk ck bi ci aij cj
◦ j k i,j
◦ ◦
◦ [[τ 2 ]] {2 f 2 }2 = f f (f, f ) 4 2 12 1 aij ajk c2k bi aij c2j
◦ j k i,j
◦
◦
◦ [[τ ]] {3 f }3 = f f f f 4 1 24 1 aij ajk akn cn bi aij ajk ck
◦ j k n i,j,k
1
φ(t) = bi φi = , ∀ r(t) ≤ p, t ∈ T, (1.53)
i
γ(t)
From Table 1.3, we then obtain the following number of orders trees (see Table 1.4).
order p 1 2 3 4 5 6 7 8 9 10
order p 1 2 3 4 5 6 7 8 9 10
Number of order conditions for Multi-stage R–K up to order 10, can be seen in the
following Table 1.4.
l
◦ ◦
l
j k
◦ ◦ j◦ ◦k k ◦+
R◦ R◦ s◦ j
i+ i + ◦+
+ i
τ31
(nolabeled 3-tree)
(labeled 3-trees) 1 2 3 3 1 2 2 3 1
A B C
(rooted labeled 3-tree)
(+)1-2-3 1-(+)2-3 1-2-(+)3 (+)3-1-2 3-(+)1-2 3-1-(+)2 (+)2-3-1 2-(+)3-1 2-3-(+)1
a b c d e f g h i
◦ ◦ ◦
◦
(rooted 3-tree)
◦
◦
ρτ32
ρτ31
Fig. 1.5. The 3-tree, labeled 3-trees (A) − (C), rooted labeled 3-tree (a) − (i), and rooted
3-tree
7.1 Multistage Symplectic Runge–Kutta Method 299
•
ρτ11 + τ11 •
•
ρτ21 • τ21 ••
+
• • •
ρ τ31 • ρτ32 τ31 • • •
•
• +
+
• •
ρτ41 • ρτ42 • • τ41 • • • •
• 1 2 3 4
• •
+ +
•
• •
ρτ43
•
ρτ44 • • • τ42 • •
• • •
+ +
o p
m n • •
k l • •
• •
j •
+ •
i ρτ i
• • • •
• • • •
• • • •
• • • •
+ + ρτj +
ρτI ρτJ
Superfluous trees. Let τ be an n-tree and choose one of its labelings λτ . This
labeling gives rise to n different rooted labeled trees ρλτ1 , · · · , ρλτn , where ρλτi has
its root at the integer i (1 ≤ i ≤ n). If for each edge (i, j) in λτ , ρτi and ρτj represent
different rooted trees; then τ is called non-superfluous. Consider the 3-tree τ31 in
300 7. Symplectic Runge–Kutta Methods
Fig.1.5. When choosing the labeled 3-tree A, we see that for the edge 1-2, choosing 1
as the root leads to ρτ31 , and choosing 2 as the root leads to ρτ32 . For the edge 2-3,
choosing 2 as the root leads to ρτ32 , and choosing 2 as the root leads to ρτ31 . Therefore
τ31 is non-superfluous. One the other hand the 4-tree with labeling is superfluous (see
Fig. 1.6 and 1.7), since changing the root from 2 to the adjacent 3 does not result in
different rooted trees. • • • •
1 2 3 4
In order to simplify the order conditions for symplectic R–K,we need some lem-
mas. Before introducing the lemmas, let us first look at Fig. 1.8: 4-rooted tree.
Look at first rooted tree ρτi (i.e., root at i) and rooted tree ρτj . The root of the
rooted trees ρτI , ρτJ in Fig. 1.8 is at vertex i and j, they are removed edge joining i
and j in the top left-hand corner graph.
1 1
φ(ρτi ) + φ(ρτj ) = + . (1.56)
γ(ρτi ) γ(ρτj )
Therefore ρτi order conditions hold iff order conditions of ρτj hold.
γ(ρτI )
γ(ρτi ) = rγ(ρτJ ) , (1.57)
r(ρτI )
γ(ρτJ )
γ(ρτj ) = rγ(ρτI ) , (1.58)
r(ρτJ )
where r(ρτI ) and r(ρτJ ) are the orders of ρτI and ρτJ . Then, insert r(ρτI ) in formula
(1.57) and r(ρτJ ) in formula (1.58) into r(ρτI )+r(ρτJ ) = r to obtain (1.54). Rewrite
the left side of formula (1.55) into
; ;
φ(ρτi ) + φ(ρτj ) = bi aij + bj aij , (1.59)
ij··· ij···
B
where represents a product of r − 2 factors akl . Equality (1.55) can be obtained
using order condition (1.15) of the symplectic R–K method.
Example 1.26. See the simple examples below. From Fig. 1.8 we have
7.1 Multistage Symplectic Runge–Kutta Method 301
φ(ρτ v ) = biv aiv iw aiv i1 aiv i2 aiw i3 aiw i4 ,
1 2 •4
vw1···4 • • •3
φ(ρτ w ) = biw aiw iv aiw i3 aiw i4 aiv i1 aiv i2 . • •
v w
vw1···4
1
φ(ρτ ) = . (1.60)
γ(ρτ )
Proof. Choose first a non-superfluous tree τ . Assume that condition (1.60) is satisfied
for a suitable rooted tree ρτi of τ . From the Lemma 1.25 we choose j as any of the
vertices adjacent to i. By condition (1.56), the order condition (1.60) is also satisfied
for ρτj . Since any two vertices of a tree can be joined through a chain of pairwise
adjacent vertices, the iteration of this argument leads to the conclusion that the method
satisfies the order conditions that arise from any rooted tree in τ . In the case of a
superfluous tree τ , by definition, it is possible to choose adjacent vertices i, j, such
that ρτi and ρτj are in fact the same rooted tree. Then condition (1.56) shows that
(1.60) holds for the rooted tree ρτi . Therefore (1.60) holds for all rooted tree in τ .
Example 1.28. For r = 2, there is only one tree τ21 , this is a superfluous tree.
Example 1.29. For r = 3, there is again only one tree τ31 . It has two rooted trees
ρτ31 , ρτ32 . Hence the order conditions become
3 3
1 1
bi aij aik = , or bi aij ajk = .
3 6
i,j,k=1 i,j,k=1
Example 1.30. For r = 4, there is only one non-superfluous tree τ42 . We impose
either the order conditions for ρτ43 or the order conditions for ρτ44 .
We see that for symplectic R–K methods it is sufficient to obtain the order con-
ditions only for non-superfluous trees rather than every rooted trees. The reduction
in the number of order conditions is given in Table 1.5. Comparison order conditions
between symplectic R–K (R–K–N) method.
302 7. Symplectic Runge–Kutta Methods
Example 1.31. For diagonally symplectic R–K method, see tableau (1.16).
If r = 3, according to Theorem 1.27 and Table 1.5, symplectic R–K method has
only two conditions. One condition is for r = 1,
b1 + b2 + b3 = 1. (1.61)
another condition is for r = 3, which has only one non-superfluous tree with two
rooted trees, ρτ31 , ρτ32 . Choose one of them
3
1
r(ρτ31 )φ(ρτ31 ) = 1, bi c2i = ,
i=1
3
Let us suppose that the component p of the first set of system (2.2) are integrated by
an R–K method and the component q in second part of system are integrated with
a different R–K method. The overall scheme is called a Partitioned– Runge–Kutta
method, or shortly called P–R–K method. It can be specified by two Butcher tableaux:
Proof. Let
Ki = f (Qi ), li = g(Pi ), (2.6)
and
d pn+1 ∧ d q n+1 − d pn ∧ d q n
s
= h bi d Ki ∧ d Qi + Bi d Pi ∧ d li
i=1
s
−h2 (bi Aij + Bj aji − bi Bj )d Ki ∧ d lj . (2.7)
i,j=1
Note that the first term on the right side of Equation (2.7) is
304 7. Symplectic Runge–Kutta Methods
bi d Ki ∧ d Qi + Bi d Pi ∧ d li
s
= −vqq d Qi ∧ d Qj + upp d Pi ∧ d Pj = 0.
i,j=1
bi Aij + Bj aji − bi Bj = 0.
introducing matrix
⎡ ⎤
p0 (c1 ) p1 (c1 ) · · · ps−1 (c1 )
⎢ ⎥
⎢ p0 (c2 ) p1 (c2 ) · · · ps−1 (c2 ) ⎥
⎢ ⎥
W =⎢ ⎥,
⎢ ··· ··· ··· ··· ⎥
⎣ ⎦
p0 (cs ) p1 (cs ) · · · ps−1 (cs )
W T BW = I.
√
k k k + i
pk (x) = 2k + 1 (−1)k+i xi , k = 0, 1, · · · , s − 1.
i i
i=0
For an s-stage R–K method (A, b, c), let X = W −1 AW = W T BAW , then for the
high order R–K method based on the high order quadrature formula their transforma-
tion matrix X is given by Table 2.1[Sun93a,Sun94,Sun95] .
7.2 Symplectic P–R–K Method 305
1
0 1 0 0
=⇒ 2
1 1 1
306 7. Symplectic Runge–Kutta Methods
1 1 1 1
0 − 0 0 0 −
4 4 8 8
2 1 5 2 1 1 =⇒ 7 3
3 4 12 3 3 3 24 8
1 3 1 3 1 3
4 4 4 4 4 4
Example 2.8. P–R–K method with first order accuracy Radau II A – Radau II A
1
1 1 1 0
=⇒ 2
1 1 1
Example 2.9. P–R–K method with third order accuracy Radau II A – Radau II A
1 5 1 1 1 3 1
− 0 −
3 12 12 3 3 8 24
3 1 =⇒ 7 1
1 1 1 0
4 4 8 8
3 1 3 1 3 1
4 4 4 4 4 4
Corollary 2.10. Using same method constructed symplectic P–R–K method Lobatto
III C – III C with 2s − 2 order accuracy.
Example 2.11. Symplectic P–R–K Lobatto III C–III C method with 2 order accuracy,
its coefficients are
1 1 1 1
0 − 0 0 0 −
2 2 4 4
1 1 =⇒ 3 1
1 1 1 0
2 2 4 4
1 1 1 1 1 1
2 2 2 2 2 2
Example 2.12. Symplectic P–R–K Lobatto III C – III C method, its coefficients are
7.2 Symplectic P–R–K Method 307
1 1 1 1 1 1
0 − 0 0 0 0 −
6 3 6 12 6 12
1 1 5 1 1 1 1 5 4 1
− 0 −
2 6 12 12 2 4 4 =⇒ 24 12 24
1 2 1 1 5 1
1 1 0 1 0
6 3 6 12 6 12
1 2 1 1 2 1 1 2 1
6 3 6 6 3 6 6 3 6
Corollary 2.13. The s-stage P–R–K Lobatto III A–III B method is symplectic with
2s − 2 order accuracy.
Example 2.14. Symplectic P–R–K Lobatto III A–III B method, its coefficients are
1 1
0 0 0 0 0 0
2 4
1 1 1 =⇒ 1 1
1 1 0
2 2 2 2 4
1 1 1 1 1 1
2 2 2 2 2 2
1 1 1 1
0 0 0 − 0 − 0
6 6 12 12
5 1 1 1 1 2 1 1
− 0 −
24 3 24 6 3 =⇒ 16 3 48
1 2 1 1 5 1 2 1
0
6 3 6 6 6 6 12 12
1 2 1 1 2 1 1 2 1
6 3 6 6 3 6 6 3 6
With the help of symplectic conditions of P–R–K methods, we can construct sym-
plectic R–K method. We have the following corollary:
1
Corollary 2.15. [Sun][Sun00] The s-stage R–K method with coefficients a∗ij = (aij +
2
Aij ), b∗i = bi = Bi and c∗i = ci are symplectic and at least satisfy B(p), C(ξ) and
D(ξ), i.e., order
r = min(p, 2ξ + 1), where ξ = min (η, ζ).
Example 2.16. If we take the coefficients in Example 2.14, we know that the right
of
the table is a special case of 2-order accuracy of R–K methods of Lobatto III S i.e.,
1
σ = situation of literature [Chi97] , see Table 2.1 .
2
pi = pi−1 + ci hf (qi−1 ),
qi = qi−1 + di hg(pi ), i = 1, 2, · · · , s, (2.8)
where f = −vq (q), g = up (p). We can regard f, g as a function of z = (p, q), for f
(or g) with the p (or q) variables of coefficient 0, i.e., f (q, 0 · p) (or g(p, 0 · q). In order
to facilitate the writing in a unified form, we make:
p = ya , q = yb , f = fa , g = fb , ya,0 = p0 , yb,0 = q0 ,
and ya,1 = ps−1 , yb,1 = qs−1 , then Equation (2.8) is transformed into an s-stage
P–R–K form:
g1,a = ya,0 = p0 ,
g1,b = yb,0 = q0 ,
g2,a = ya,0 + c1 τ fa (q0 ) = ya,0 + c1 hfa (g1,b ) = p1 ,
g2,b = yb,0 + d1 τ fb (p1 ) = yb,0 + d1 hfb (g2,a ) = q1 ,
..
. (2.9)
s−1
gs,a = ya,0 + h cj fa (gj,b ) = ps−1 ,
j=1
s−1
gs,b = yb,0 + h dj fb (gj+1,a ) = qs−1 .
j=1
(2.9) is equivalent to
i−1
gi,a = ya,0 + h cj fa (gj,b ),
j=1
i−1
gi,b = yb,0 + h dj fb (gj+1,a ), i = 2, · · · , s,
j=1
(2.10)
s−1
ya,1 = ya,0 + h cj fa (gj,b ),
j=1
s−1
yb,1 = yb,0 + h dj fb (gj+1,a ).
j=1
Let
a1 = c1 , a2 = c2 , ···, as−1 = cs−1 , as = 0,
(2.12)
b1 = 0, b2 = d1 , ···, bs−1 = ds−2 , bs = ds−1 ,
then schemes (2.10) now become
7.2 Symplectic P–R–K Method 309
i−1
i−1
gi,a = ya,0 + h aj fa (gj,b ) = ya,0 + h aj Rj,a ,
j=1 j=1
i
i
gi,b = yb,0 + h dj fb (gj,a ) = yb,0 + bj Rj,b , i = 2, · · · , s,
j=1 j=1
(2.13)
s
ya,1 = ya,0 + h aj Rj,a ,
j=1
s
yb,1 = yb,0 + h bj Rj,b .
j=1
Where
Ri,a = fa (gi,b ), Ri,b = fb (gi,a ). (2.14)
Now, we just need to study the order conditions of scheme (2.13) when as = b1 = 0.
Notice that as = b1 = 0 is necessary for (2.13) to be canonical and is also crucial for
simplifying order conditions, as we will see later.
A P -graph (denoted by P G) is a special graph which satisfies the following con-
ditions: (i) its vertices are divided into two classes: “white” ◦ “black” •, sometimes
instead “meagre” and “fat”. (ii) the two adjacent vertices of a P G cannot be of the
same class. If we give the vertices of PG an arbitrary set of labels, we get a label P -
graph, and we say P -graph∈ P G. Two labeled P -graphs are said to be isomorphic
labeled P -graphs if they are just two different labelings of the same P -graph.
A simple path joins a pair of vertices v and w, v = w, and is a sequence of pairwise
distinct vertices v = v0 , v1 , · · · , vn−1 = w, where vi = vi−1 , (vi−1 , vi ) ∈ E. Fig.
2.1 shows an example of a simple path of v and w for n = 4.
◦ ◦ ◦
Fig. 2.1. A simple path of v and w black • and white ◦, for n=4
K(χ(v1 )) = K(v1 ),
where v1 ∈ V1 , and
1, for v black,
K(v) =
0, for v white.
A P -tree (n order) is the equivalent of such a class, it consists of a labeled P -tree and
all of its isomorphism. We use [HNW93] the P -series and tree method to derive the order
condition of Equation (2.13) below. We first introduce some definitions and notations.
(3) Two rooted labeled P -trees with same order, {V1 , E1 , r1 } and {V2 , E2 , r2 }
(where ri (i = 1, 2) denoted rooted label), are called rooted isomorphism if there
exists a χ that satisfies the condition of (2), and χ(r1 ) = r2 holds.
A rooted P -tree, denoted by ρP τ , is an equivalence class which contains a labeled
P -tree and all of its isomorphic P -trees. We denote ρP τa (ρP τb ) as ρP τ with white
(black) root, and ρλP τ as rooted labeled P -tree, which is obtained by adding label to
ρP τ . Thus ρλP τ ∈ ρP τ .
We denote by ρP τa (resp. ρP τb ) for a rooted P -tree ρP τ that has a white (resp.
black) root. If we give the vertices of a rooted P -tree ρP τ such a set of labels so
that the label of a father vertex is always smaller than that of its sons, we then get a
monotonically labeled rooted P -tree M ρλP τ . We denote by α(ρP τ ) the number of
possible different monotonic labelings of ρP τ when the labels are chosen from the set
Aq = { the first q letters of i < j < k < l < · · ·}, where q is the order of ρP τ .
The set of all rooted trees of order n with a white (resp. black) root is denoted by
T Pna (resp. T Pnb ). Let us denote by λP τna (resp. λP τnb ) the set of all rooted labeled
P -trees of order n with a white (resp. black) root vertex, and M λP τna (resp. M λP τnb )
the set of all monotonically labeled P -tree of order n with a white (resp. black) root
vertex when the labels are chose from the set An .
(4) The density γ(ρλτ ) of a rooted P -tree ρλτ is defined recursively as
where r(ρλτ ) is order of ρλτ , and ρλτ 1 , · · · , ρλτ m are the sub-trees which arise when
the root of ρλτ is moved from the tree. The density of rooted P -tree ρP τ is calculated
by regarding them as general rooted tree neglecting the difference between with the
black and white vertices .
(5) Let ρP τ 1 , · · · , ρP τ m be rooted P -tree. We denote by ρP τ = a [ρP τ 1 , · · · ,
ρP τ m ] the unique rooted P -tree that arises when the roots of ρP τ 1 , · · · , ρP τ m are
all attached to a white root vertex. Similarly denote it by b [ρP τ 1 , · · · , ρP τ m ] when
the root of the P -tree is black. We say ρP τ 1 , · · · , ρP τ m are sub-trees of ρP τ . We
further denote the rooted P -tree of order 1, which has a white (resp.black) root vertex
by ta (resp.tb ).
(6) (2.11) is defined recursively as:
We see that F (ρP τ ) is independent of labeling. Here, and in the remainder of this
book, in order to avoid sums and unnecessary indices, we assume that ya and yb are
scalar quantities , and fa ,fb scalar functions. All subsequent formulas remain valid for
vectors if the derivatives are interpreted as multi-linear mapping.
(2.16)
where q = 1, 2, · · ·.
It is convenient to introduce two new rooted P -trees of order 0: ∅a and ∅b . The
corresponding elementary differential are F (∅a ) = ya , F (∅b ) = yb .
We further set
T P a = ∅a ∪ T P1a ∪ T P2a ∪ · · · ,
T P b = ∅b ∪ T P1b ∪ T P2b ∪ · · · ,
λT P a = ∅a ∪ λT P1a ∪ λT P2a ∪ · · · , (2.17)
λT P b = ∅b ∪ λT P1b ∪ λT P2b ∪ · · · ,
M λT P a = ∅a ∪ M λT P1a ∪ M λT P2a ∪ · · · ,
M λT P b = ∅b ∪ M λT P1b ∪ M λT P2b ∪ · · · .
p-series: let c(∅a ), c(∅b ), c(ta ), c(tb ), · · · be real coefficients defined for all P -
trees, c : T P a ∪ T P b → R. The series p(c, y) = (pa (c, y), pb (c, y)) is defined
as
hr(ρλP τ )
pa (c, y) = c(ρλP τ )F (ρλP τ )(y)
r(ρλP τ )!
ρλP τ ∈M λT P a
hr(ρλP τ )
= α(ρP τ ) c(ρλP τ )F (ρλP τ )(y),
(ρλP τ )!
ρP τ ∈T P a
(2.18)
hr(ρλP τ )
pb (c, y) = c(ρλP τ )F (ρλP τ )(y)
(ρλP τ )!
ρλP τ ∈M λT P b
hr(ρλP τ )
= α(ρP τ ) c(ρλP τ )F (ρλP τ )(y).
r(ρλP τ ) !
ρP τ ∈T P b
312 7. Symplectic Runge–Kutta Methods
where Y (ρP τ ) = 1 for all rooted P -trees ρP τ . The following theorem is from the
book[HNW93] .
with
Let
⎧
⎪
⎪ Ri,a = pa Ki , (ya,0 , yb,0 ) ,
⎪
⎨ R = p K , (y , y ) ,
i,b b i a,0 b,0
i = 1, · · · , s, (2.19)
⎪
⎪ g = p G , (y , yb,0 ) ,
⎪
⎩
i,a a
i a,0
gi,b = pb Gi , (ya,0 , yb,0 ) ,
where Ki (i = 1, · · · , s) : T P a ∪ T P b → R, Gi (i = 1, · · · , s) : T P a ∪ T P b → R
are two sets of p-series. From (2.10), we have Gi (∅a ) = Gi (∅b ) = 1. Hence, From
(2.14), we have
pa Ki , (ya,0 , yb,0 ) ! Ri,a
!
fa (gi,b )
!
p Ki , (ya,0 , yb,0 ) = = =h
pb Ki , (ya,0 , yb,0 ) Ri,b fb (gi,a )
fa pb Gi , (ya,0 , yb,0 ) ! fa p Gi , (ya,0 , yb,0 ) !
= h = h
fb pa Gi , (ya,0 , yb,0 ) fb p Gi , (ya,0 , yb,0 )
= p Gi , (ya,0 , yb,0 ) .
Ki = Gi , ∀ i = 1, · · · , s.
⎡ ⎤
i−1
⎢ ya,0 + h aj Rj,a ⎥
pa Gi , (ya,0 , yb,0 ) ! ⎢ ⎥
⎢ j=1 ⎥
p Gi , (ya,0 , yb,0 ) = =⎢ ⎥
pb Gi , (ya,0 , yb,0 ) ⎢ i ⎥
⎣ yb,0 + h bj Rj,b ⎦
⎡ ⎤j=1
i−1
⎢ ya,0 + h aj pa Kj , (ya,0 , yb,0 ) ⎥
⎢ ⎥
⎢ j=1 ⎥
= ⎢
⎥
.
⎢ i ⎥
⎣ yb,0 + h bj pb Kj , (ya,0 , yb,0 ) ⎦
j=1
Thus:
i−1
Gi (ρP τa ) = aj Kj (ρP τa ),
j=1
∀ r(ρP τ ) ≥ 1. (2.20)
i
Gi (ρP τb ) = bj Kj (ρP τb ),
j=1
Comparing the numerical solution obtained from (2.13) and the exact solution from
(2.11), we get the order condition for scheme (2.13).
Theorem 2.19. Scheme (2.13) is p-order accuracy iff its coefficients ai , bi satisfy:
⎧ s
⎪
⎪
⎪ ai Ki (ρP τa ) = 1, 1 ≤ r(ρP τa ) ≤ p,
⎨
i=1
(2.22)
⎪
⎪
s
⎪
⎩ b i K i (ρP τ b ) = 1, 1 ≤ r(ρP τ b ) ≤ p,
i=1
From the first and second equations of (2.23) we know Ki (∅a ) = Ki (∅b ) =
0, Ki (ta ) = Ki (tb ) = 1, from the last two equations of (2.23) we can obtain
Gi (ta ), Gi (tb ). Repeating this procedure, we can obtain Ki (ρP τa ), Ki (ρP τb ), by P -
tree order from low to high.
314 7. Symplectic Runge–Kutta Methods
Next we rewrite equations (2.23) into more intuitive forms. From (2.23), we have
⎧ * i + * i +
⎪
⎪
⎪
⎪ Ki (ρP τa ) = r(ρP τa ) 1
bj Kj (ρP τb ) · · · m1
bj Kj (ρP τb ) ,
⎪
⎨ j=1 j=1
* i−1 + * i−1 + i = 2, 3, · · · , s,
⎪
⎪
⎪
⎪ Ki (ρP τb ) = r(ρP τb ) 1
aj Kj (ρP τa ) · · · m 2
⎪
⎩ aj Kj (ρP τa ) ,
j=1 j=1
(2.24)
where
ρP τa = a [ρP τb1 , · · · , ρP τbm1 ],
(2.25)
ρP τb = b [ρP τa1 , · · · , ρP τam2 ].
We now define elementary weight φ(ρP τ ) for a rooted P -tree. Choose any labeling
ρλP τ for ρP τ ; without loss of generality we choose a monotonic one with labels
i < j < k < l < · · ·, where the rooted labeling is i. Then φ can be obtained
recursively (Note the difference of solving for φ between the original tree and its sub-
tree)
⎧
⎪
s−1
⎪
⎪ ai φ(ρP τb1 ), · · · , φ(ρP τbm1 ) ,
⎪
⎪
φ(ρP τ a ) =
⎪
⎨ i=1
s
r(ρP τa ), r(ρP τb ) ≥ 1,
⎪
⎪ φ(ρP τ ) = bi φ(ρP τa1 ), · · · , φ(ρP τam2 ) ,
⎪
⎪
b
⎪
⎪ i=1
⎩
φ(∅a ) = φ(∅b ) = 1,
(2.26)
where ρP τa , ρP τb and (2.25) are the same. Here, notice that i is the root of ρP τa or
ρP τb , and s is the label for an imaginary father vertex to the root i. The summation is
always with respect to subscripts of the son’s vertex, from 1 adds to the father vertex
or it reduces by 1. Now s is order of scheme (2.13). We are doing this only for ease
of the recursive definition; otherwise vertex i has no father vertex and the summation
superscript cannot be determined. Regarding the subtrees of ρP τa , ρP τb , the father
vertex for their root labeling is i. It is not necessary to add an extra father vertex. So
the weight for a p-tree as the original tree is different from the weight as another subset
tree. By (2.26) we can see that the elementary weight of a tree φ is not related to its
labeling as long as the imaginary father of maintaining root label is always the order
of the scheme (2.13).
[AS93,ZQ95b,SSC94]
Theorem 2.20. Order conditions in Theorem 2.19 are equivalent to
1
φ(ρP τ ) = , ∀ ρP τ ∈ T P a ∪ T P b , r(ρP τ ) ≤ p. (2.27)
γ(ρP τ )
Proof. We just need to prove
⎧
⎪
s−1
⎪
⎪
⎪
⎨
φ(ρP λτa )γ(ρλP τb ) = aj Kj (ρP τa ),
j=1
(2.28)
⎪
⎪
s
⎪
⎪ φ(ρλP τ )γ(ρλP τ ) = bj Kj (ρλP τb ).
⎩ b a
j=1
7.2 Symplectic P–R–K Method 315
so we have to prove
i
φ(ρλP τbl )γ(ρλP τbl ) = bjn Kjn (ρP τbl ) for n = 1, · · · , m1 ,
jn =1
and
i−1
φ(ρP τal )γ(ρP τal ) = ajn Kjn (ρP τal ) for n = 1, · · · , m2 .
jn =1
where l is the label of ta or tb and f (l) is the label of the father. Since
f (l)−1 f (l)−1 f (l)−1
φ(ta )γ(ta ) = al · 1 = al · Kl (ta ) = al ,
l=1 l=1 l=1
f (l) f (l) f (l)
φ(tb )γ(tb ) = bl · 1 = bl · Kl (tb ) = bl ,
l=1 l=1 l=1
and
Kl (ta ) = 1, Kl (tb ) = 1.
The theorem is proved.
316 7. Symplectic Runge–Kutta Methods
i.e., 1◦ .
Also has
s−1 ;
iv
iv ;
iw
φ(ρP τ v ) = aiv b iw ,
iv =1 1 iw =1 2
s ; w −1
iw i ;
iv
φ(ρP τ w ) = b iw aiv ,
iw =1 2 iv =1 1
iv
; iw
;
where resp. is the product of all φ(ρP τ v ) resp. φ(ρP τ w ) , while
1 2
ρλP τa = a [ρP τb1 , · · · , ρλP τbm1 ], ρλP τb = b [ρλP τa1 , · · · , ρλP τam2 ],
iv
; iw
;
and iv , iw are labels of v and w respectively. resp. varies only according to
1 2
iv , (resp. iw ), therefore
s−1 ;
iv
s ;
iw
φ(ρP τv ) = aiv , φ(ρP τw ) = biw ,
iv =1 1 iw =1 2
7.2 Symplectic P–R–K Method 317
then
s−1 ;
iv
s ;
iw
φ(ρP τv )φ(ρP τw ) = aiv bi w
iv =1 1 iw =1 2
s−1 ;
i v iv ;
iw
s iw
;
= aiv biw + bi w
iv =1 1 iw =1 2 iw =iv +1 2
s−1 ;
iv
iv ;
iw
s−1 ;
iv
s ;
iw
= aiv bi w + aiv bi w .
iv =1 1 iw =1 2 iv =1 1 iw =iv +1 2
s−1 ;
iv
s ;
iw
s ; w −1
iw i ;
iv
aiv bi w = biw aiv
iv =1 1 iw =iv +1 2 iw =2 2 iv =1 1
s ; w −1
iw i ;
iv
= bi w aiv , b1 = 0.
iw =1 2 iv =1 1
Proof. By Corollary 2.22, we know that any two kinds of rooted label method of the
P τ lead to equivalent conditions. Therefore, we only need to take one of them to get
the order conditions.
• •
• • •
τa τb a[τb ] b[τa , τa ] b[τb , τb ]
• •
• •
• •
b[a [τb , τb ]] a[b [τa , τa ]] a[a [a [τb ]]]
Finally, according to Theorem 2.23, we can simplify the order conditions for
P–R–K method, which is given in Table 2.2. Calvo and Hairer[CH95] further reduce
the number of independent condition in P–R–K method. See Table 2.3. For general
Hamiltonian, the corresponding values are given by Table 2.4 which is obtained by
Mirua[Mur97] .
Table 2.2. Order conditions P–R–K method and Symplectic P–R–K method for separable
case
Table 2.3. Further reduction in Order conditions for P–R–K method in separable case
Order P–R–K method Symp. P–R–K method expl. Symp. P–R–K method
1 2 1 1
2 2 1 1
3 4 2 2
4 8 3 3
5 18 6 6
6 40 10 9
7 96 22 18
8 230 42 30
Table 2.4. Order conditions P–R–K method and Symplectic P–R–K method for general case
ÿ J = f J (y 1 , y 2 , · · · , y n ), J = 1, · · · , n, (y 1 , · · · , y n ) ∈ Rn . (3.1)
! !
ẏ ẏ
= , (3.2)
ÿ f (y)
bj = bj (1 − cj ), 1 ≤ j ≤ s, (3.5)
bi aij − bj aji + bi bj − bi bj = 0, 1 ≤ i, j ≤ s, (3.6)
Proof. The proof of Theorem 3.1 can be find in [Sur89,OS92] . Here, we only point out
that under conditions (3.5), (3.6) are equivalent to
bi aij − bj aji + bi bj (cj − ci ) = 0, 1 ≤ i, j ≤ s. (3.7)
Therefore, the theorem is completed.
Similar to Section 7.1, we first introduce some necessary definitions and notations,
and then derive order conditions. Some definitions of Section 7.1 can still be used.
Here we only introduce some special definitions and notations.
(1) S-graph. A S-graph, denoted as S-g, is a special P -graph where any two
adjacent vertices belong to different categories: “ white (meagre)” or “ black (fat)”.
The Labeled S-graph has a definition similar to the labeled P -graph.
(2) S-tree. A S-tree, denoted as Sτ , has a definition similar to the P -tree of
7.2: replacing the P -graph in the definition of original P -tree with S-graph gives the
definition of S-tree. The definition of labeled S- tree, λSτ , rooted S-tree, ρSτ , rooted
labeled S-tree, ρλSτ , and isomorphic labeled S-trees, root-isomorphic labeled S-tree
are defined using the same method as we have used to define P -trees, labeled P -trees,
etc. We should point out that in this section, we just consider S-trees with “ black” root
vertices. So when we refer to rooted S-tree, we mean that its S-tree has a “ black” root.
Moreover, order r, density γ also has similar definition as mentioned in Section 7.1.
But the elementary weight definition is completely different, which we will redefine
subsequently.
Definition 3.2. We define the elementary weight φ(ρλSτ ) corresponding to a rooted
labeled S-tree. At first, for convenience, we assume ρλSτ is monotonically labeled.
Later, we will see this is unnecessary. In the remainder of this section, without speci-
fication, the labels of the vertices are always j < k < l < m < · · ·. For a monotonic
labeling, the label of the root is j. Then φ(ρλSτ ) is a sum over the labels of all fat
vertices of ρλsτ , the general term of the sum is a product of
1◦ bj (j is a rooted vertex).
2◦ akl , if the fat vertex k is connected via a meagre son with another fat vertex l.
3◦ cm k , if the fat vertex k has m meagre end-vertices as its sons, where an end-
vertex is the vertex which has no son.
We see that, for two different rooted labeled S-trees: ρλSτ 1 and ρλSτ 2 , we have
φ(ρλSτ 1 ) = φ(ρλSτ 2 ). Thus, the choosing of the monotonic labeling is unnecessary.
m j
k • •l m • •l
For example, for and , we have
+j +k
ρλSτ 1 ρλSτ 2
φ(ρλSτ 1 ) = bj cj ajm = bk akj ck = φ(ρλSτ 2 ) = φ(ρSτ ).
j,m j,k
1 2
Because ρλSτ and ρλSτ are rooted isomorphism, they belong to a rooted tree ρSτ :
• •
+ . Therefore, they form an equivalence class. The following theorem can be seen
in the literature [HNW93] . We omit the proof here.
322 7. Symplectic Runge–Kutta Methods
The explanation for φ (ρSτ ) is similar to that of φ(ρSτ ), which only needs to sub-
stitute bj in φ(ρSτ ) (suppose j is the label of rooted tree, corresponding to a certain
label choosing) by b̄j . Because φ and φ is independent of the chosen label, (3.8) and
(3.9) can take any of the labels to calculate.
We find that (3.8) and (3.9) are not independent under symplectic conditions.
Theorem 3.4. Under symplectic condition (3.5), order condition (3.8) implies condi-
tion (3.9)[ZQ95b] .
Proof. Let ρSτ be ≤ p − 1 order S-tree, and let ρSu be such a rooted S-tree with
r(ρSτ ) + 1 order that is obtained from ρSτ rooted tree by attaching a new branch
with a meagre vertex to the root of τ . Therefore from definition of φ, we have
; ;
φ(ρSu) = bj cj , φ(ρSτ ) = bj ,
j j
where we assume that ρSτ and ρSu have monotonic labels j < k < l < · · ·. Then for
ρSu, apart from the added root of the meagre leaf node, the remaining vertices ;have
the same labeling as ρSτ , and is a sum for all non-fat root vertices, and is a
product of aij and ci that are contained in ρSτ and ρSu. From the definition of φ, we
have
(r(ρSτ ) + 1)γ(ρSτ )
γ(ρSu) = ,
r(ρSτ )
therefore,
; ;
φ (ρSτ ) = bj = bj (1 − cj )
j j
; ;
= bj − bj cj = φ(ρSτ ) − φ(ρSu)
j j
1 1 1
= − = . (3.10)
γ(ρSτ ) γ(ρSu) (r(ρSτ ) + 1)γ(ρSτ )
Since the formula(3.8) is held for ≤ p-order S-tree and (3.9) is held only for
≤ p − 1-order S-tree, the order of the tree obtained by adding a leaf node to any
≤ p − 1-order tree on the root must be ≤ p and (3.8) must be satisfied. Therefore the
final equal sign in (3.10) holds. Thus we reach the conclusion of this section.
Note that the conditions we have given here are necessary and sufficient. How-
ever some conditions of (3.13) are still redundant, which means some conditions are
mutually equivalent. We will see more details about this in Section 7.4.
For conditions of 3-stage of 3rd-order R–K–N method, we get equations for parame-
ters : ⎧
⎪ b2 a21 + b2 b1 (c1 − c2 ) = 0,
⎪
⎪
⎪
⎪
⎪ b3 a31 + b3 b1 (c1 − c3 ) = 0,
⎪
⎪
⎪
⎪
⎪ b3 a32 + b3 b2 (c2 − c3 ) = 0,
⎪
⎪
⎨ b1 + b2 + b3 = 1,
(3.14)
⎪
⎪ b c + b c + b c =
1
,
⎪
⎪ 1 1 2 2 3 3
2
⎪
⎪
⎪
⎪ 2 2 2 1
⎪
⎪ b1 c1 + b2 c2 + b3 c3 = ,
⎪
⎪ 3
⎪
⎩ 1
b2 a21 + b3 a31 + b3 a32 = ,
6
and
bi = bi (1 − ci ), i = 1, 2, 3. (3.15)
324 7. Symplectic Runge–Kutta Methods
b2 a21 + b2 b1 (c1 − c2 ) = 0,
b3 a31 + b3 b1 (c1 − c3 ) = 0,
b3 a32 + b3 b2 (c2 − c3 ) = 0,
b4 a41 + b4 b1 (c1 − c4 ) = 0,
b4 a42 + b4 b2 (c2 − c4 ) = 0,
b4 a43 + b4 b3 (c3 − c4 ) = 0,
b1 + b2 + b3 + b4 = 1,
1
b1 c1 + b2 c2 + b3 c3 + b4 c4 = ,
2
2 2 2 2 1
b1 c1 + b2 c2 + b3 c3 + b4 c4 = ,
3
1
b2 a21 + b3 a31 + b3 a32 + b4 a41 + b4 a42 + b4 a43 = ,
6
1
b1 c31 + b2 c32 + b3 c33 + b4 c34 = ,
4
1
b2 c2 a21 + b3 c3 a31 + b3 c3 a32 + b4 c4 a41 + b4 c4 a42 + b4 c4 a43 = ,
8
1
b2 a21 c1 + b3 a31 c1 + b3 a32 c2 + b4 a41 c1 + b4 a42 c2 + b4 a43 c3 = .
24
Set c4 = 0, we have
b4 a42 + b4 b2 c2 = 0, (3.20)
b4 a43 + b4 b3 c3 = 0, (3.21)
b1 + b2 + b3 + b4 = 1, (3.22)
1
b1 c1 + b2 c2 + b3 c3 = , (3.23)
2
1
b1 c21 + b2 c22 + b3 c23 = , (3.24)
3
1
b2 a21 + b3 a31 + b3 a32 + b4 a41 + b4 a42 + b4 a43 = , (3.25)
6
1
b1 c31 + b2 c32 + b3 c33 = , (3.26)
4
1
b2 c2 a21 + b3 c3 a31 + b3 c3 a32 = , (3.27)
8
1
b2 a21 c1 + b3 a31 c1 + b3 a32 c2 + b4 a41 c1 + b4 a42 c2 + b4 a43 c3 = . (3.28)
24
We obtain a set of numerical solutions of (3.16) – (3.28):
a21 = 0.2232896E − 01, a31 = 0.2822977E − 08,
a32 = 0.2886753, a41 = 0.3053789E − 01,
a42 = −0.1057342, a43 = −0.4251137,
b1 = −0.3867491E − 01, b2 = 0.5000003,
b3 = 0.5386746, b4 = −0.9129767E − 07,
c1 = 0.7886753, c2 = 0.2113249,
c3 = 0.7886752.
Guessing from these numerical solutions, we obtain
1 1
a31 = 0, b2 = , b3 = − b1 , b4 = 0, c1 = c3 , c2 = 1 − c1 .
2 2
Inserting them into (3.16) – (3.28), we have
(3.16) ⇐⇒ a21 + b1 (2c1 − 1) = 0,
(3.17) ⇐⇒ 0 = 0,
1 1 1
(3.18) ⇐⇒ − b1 a32 + − b1 (1 − 2c1 ) = 0,
2 2 2
(3.19), (3.20), (3.21), (3.22), (3.23) ⇐⇒ 0 = 0, or 1 = 1,
1 2 1 1
(3.24) ⇐⇒ c + (1 − c1 )2 = ,
2 1 2 3
1 1 1
(3.25) ⇐⇒ a21 + − b1 a32 = ,
2 2 6
2 1
(3.26) ⇐⇒ c1 − c1 + = 0,
6
1 1 1
(3.27) ⇐⇒ (1 − c1 )a21 + c1 a32 − b1 = ,
2 2 8
1 1 1
(3.28) ⇐⇒ a21 c1 + − b1 a32 (1 − c1 ) = .
2 2 24
326 7. Symplectic Runge–Kutta Methods
√ √
2+ 3 3
Solution 2: a21 = , a31 = 0, a32 = − , a41 , a42 , a32 arbitrary.
12 6
√ √
3+2 3 1 3−2 3
b1 = , b2 = , b3 = , b4 = 0,
12 2 12
√ √ √
3− 3 3+ 3 3− 3
c1 = , c2 = , c3 = , c4 = 0.
6 6 6
Scheme 1:
ci aij
√
3+ 3
0 0 0
6
√ √
3− 3 2− 3
0 0
6 12
√ √
3+ 3 3
0 0
6 6
√ √ √
5−3 3 3+ 3 1+ 3
bi
24 12 24
√ √
3−2 3 1 3+2 3
bi
12 2 12
Scheme 2:
ci aij
√
3− 3
0 0 0
6
√ √
3+ 3 2+ 3
0 0
6 12
√ √
3− 3 3
0 − 0
6 6
√ √ √
5+3 3 3− 3 1− 3
bi
24 12 24
√ √
3+2 3 1 3−2 3
bi
12 2 12
Remark 3.6. We can obtain the required solutions easily by solving only the first 6
equations from the simplified order conditions of symplectic R–K–N.
Let Sτ be an S-tree of order n ≥ 3. It has at least two fat vertices. Let λSτ be a
labeling. Let v and w be two fat vertices connected via a meagre vertex u. For order
≤ 2 S-tree, the root S-trees contains only one the first-order and one the second-
order. Therefore there are no such issues that the order conditions for the trees with
the same order are related to each other. We consider six rooted S-trees. Let us denote
ρSτ v (resp. ρSτ w ) as the rooted S-tree obtained by regarding the vertex v (resp. w)
as the root of Sτ . Let us denote ρSτ vu (resp. ρSτ wu ) as the rooted S-tree with root
v (resp. w) that arises when the edge (u, w)(resp. (v, u)) is deleted from sτ . Let us
denote ρSτv and ρSτw as the rooted S-tree with root v and w respectively which arise
when edges, (u, v) and (u, w) are deleted from Sτ . Fig. 3.1 shows the rooted trees of
Theorem 3.7.
• + • • +
v u w v u w v u w
Sτ ρSτ v ρSτ w
+ • • + + v w +
v u u w
vu wu
ρSτ ρSτ ρSτv ρSτw
the general tree τ , with the difference between the black and white vertices neglected.
Then from (3.37), we have
% &
1 1 n 2 (y+ 1)γ(ρSτv ) − 1 (x + 1)γ(ρSτw )
− = . (3.38)
γ(ρSτ v ) γ(ρSτ w ) n 1 (x + 1)γ(ρSτv )2 (y + 1)γ(ρSτw )
Because
; ;
γ(ρSτ vu ) = (x + 1) , γ(ρSτ wu ) = (y + 1) ,
1 2
and
; ;
γ(ρSτv ) = x , γ(ρSτw ) = y ,
1 2
we have
⎛; ; ⎞
(y + 1)γ(ρSτv ) − (x + 1)γ(ρSτw )
1 1 n⎜ 2 1 ⎟
− = ⎝ ⎠
γ(ρSτ v ) γ(ρSτ w ) n γ(ρSτ vu )γ(ρSτ wu )γ(ρSτv )γ(ρSτw )
⎛ ;; 2 ⎞
(x − y 2 + x − y)
1⎜ 1 2 ⎟
= ⎝ ⎠.
n γ(ρSτ vu )γ(ρSτ wu )γ(ρSτv )γ(ρSτw )
However,
1 1
−
γ(ρSτ vu )γ(ρSτw ) γ(ρSτ wu )γ(ρSτv )
n γ(ρSτ wu )γ(ρSτv ) − γ(ρSτ vu )γ(ρSτw )
=
n γ(ρSτ vu )γ(ρSτw )γ(ρSτ wu )γ(ρSτv )
* +
; ; ; ;
n (y + 1) x− (x + 1) y
= 2 1 1 2
n γ(ρSτ vu )γ(ρSτ wu )γ(ρSτv )γ(ρSτw )
;;
(x + y + 1) x(y + 1) − (x + 1)y
1 2
=
n γ(ρSτ vu )γ(ρSτ wu )γ(ρSτv )γ(ρSτw )
;; 2
(x − y 2 + x − y)
1 2
= . (3.39)
n γ(ρSτ vu )γ(ρSτ wu )γ(ρSτv )γ(ρSτw )
Thus, we get 1◦ .
By definition of φ, we have
330 7. Symplectic Runge–Kutta Methods
⎧
⎪ ;
v ;
v
⎪
⎪ φ(ρSτ vu
) = b c , φ(ρSτv ) = bi v ,
⎪
⎨ i v i v
iv iv
⎪
⎪ ;
w ;
w
⎪
⎪ φ(ρSτ wu
) = b c , φ(ρSτw ) = bi w ,
⎩ i w iw
iw iw
and ⎧ * v w+
⎪
⎪ ;;
⎪
⎪ v
φ(ρSτ ) = biv aiv iw ,
⎪
⎨
iv iv
* v w+ (3.40)
⎪
⎪ ;;
⎪
⎪ φ(ρSτ w ) =
⎪
⎩
biw aiw iv ,
iw iw
v ;
; w
where denotes part of φ(ρSτ v )(resp. φ(ρSτ w ), which is the sum over black
vertices of ρSτv (ρSτw ). From symplectic order condition (3.11), we have
;
w ;
v
φ(ρSτ v ) − φ(ρSτ w ) = (biv aiv iw − biw aiw iv )
iv ,iw
;
v ;
w
= biv biw (civ − ciw )
iv ,iw
;
v ;
w
= biv civ biw
iv iw
;
w ;
v
− biw ciw biv
iw iv
= φ(ρSτ vu )φ(ρSτw ) − φ(ρSτ wu )φ(ρSτv ). (3.41)
Thus, the second part 2◦ of Theorem 3.7 is held. The following corollary is trivial.
Corollary 3.8. Suppose that the R–K–N method (3.4) satisfying (3.7), has order at
least n − 1, with n ≥ 3. If ρSτ v and ρSτ w are different rooted S-trees of order n,
then the order condition is the same as given in Theorem 3.7.
1
φ(ρSτ v ) =
γ(ρSτ v )
holds, iff
1
φ(ρSτ w ) =
γ(ρSτ w )
is satisfied.
1 1
φ(ρSτ vu ) = , φ(ρSτ wu ) = ,
γ(ρSτ vu ) γ(ρSτ wu )
and
1 1
φ(ρSτv ) = , φ(ρSτw ) = ,
γ(ρSτv ) γ(ρSτw )
similarly by Theorem 3.1, corollary is proved.
Proof. By Corollary 3.8 we know that any two different methods of choosing the
corresponding root have equivalent order conditions. Hence the theorem is proved.
and
5
bj = 1, (3.44)
j=1
5
1
bj cj = , (3.45)
2
j=1
332 7. Symplectic Runge–Kutta Methods
5
1
bj c2j = , (3.46)
3
j=1
5
1
bj ajl = , (3.47)
6
j,l=1
5
1
bj c3j = , (3.48)
4
j=1
5
1
bj cj ajm = , (3.49)
8
m,j=1
5
1
bj c4j = , (3.50)
5
j=1
5
1
bj c2j ajp = , (3.51)
10
j,p=1
5
1
bj ajl ajp = , (3.52)
20
j,l,p=1
5
1
bj cj ajl cl = . (3.53)
30
j,l=1
+ j + j + j + j + j
1 2 3 4 5
m•
m p l p l
l m
k• •l • • k• •l •m k • •m k• • p
k• •p
+ j + j + j + j + j
6 7 8 9 10
Fig. 3.2. Rooted S-trees corresponding to order condition (3.44) – (3.53)
For the sake of convenience, we choose monotonic labelings for trees in Fig. 3.2.
We obtain the Equations (3.46). In the following list we provide four sets of numerical
solutions, whose laws are yet to be studied further.
7.4 Formal Energy for Symplectic R–K Method 333
i 1 2 3 4 5
bi 0.396826 −0.824374 0.204203 1.002182 0.221161
1
ci 0.961729 0.866475 0.127049 0.754358 0.229296
bi 0.221160 1.002182 0.204203 −0.824375 0.396827
2
ci 0.770703 0.245641 0.872950 0.133524 0.038270
bi −1.670799 1.221431 0.088495 0.959970 0.400902
3
ci 0.694313 0.637071 −0.020556 0.795861 0.301165
bi 0.400902 0.959969 0.088495 1.221434 −1.670802
4
ci 0.698834 0.204138 1.020556 0.362928 0.305086
Remark 3.11. R–K, P–R–K, and R–K–N methods have corresponding order condi-
tions. The order conditions for symplectic R–K, symplectic P–R–K, and symplectic
R–K–N method can be simplified using symplectic conditions. The order conditions
for order 1 to 8 have already been listed in Table 1.4. Calvo and Hairer[CH95] further
reduce the number of independent condition in R–K–N method. See Table 3.1.
Table 3.1. Order conditions R–K–N method and Symplectic R–K–N method for general case
Order R–K–N method Symplectic R–K–N method
1 1 1
2 1 1
3 2 2
4 3 2
5 6 4
6 10 5
7 20 10
8 36 14
methods for symplectic R–K method mostly use Poincaré lemma, and then use the
quadrature method. Although theoretically primary function (total differential) does
exist, obtaining the primary function through the integral is not that easy. Therefore,
we attempt to calculate the formal energy of a symplectic R–K method in a easy way
that does not need the integral and also does not need any differentiation.
ż = f (z), z ∈ Rn . (4.1)
Definition 4.1. [Hai94] Let t be a rooted tree. A partition of t into k subtrees {s1 , . . . , sk }
is a set S, consisting of k − 1 branches of t such that the trees s1 , . . . , sk are obtained
when the branches of S are removed from t. Such a partition is denoted by (t, S). We
further denote α(t, S) as the number of possible monotonic labelings of t such that
the vertices of the subtrees sj are labeled consecutively.
Example 4.2. All partitions of t = [[τ ], [τ ]], t into k subtrees with the numbers
α(t, S):
t= • a(•) = b(•)
• • •
t= a = b + b ( • )2
• • •
• • • • • • 3 •
t= a = b + b b( • ) + b( • )3
• • • 2 •
• • • •
t= • a • = a • + 3b b ( • ) + b ( • )3
• • • •
1 α(t, S)
r(t)
r(t)
a(t) = b(s1 ) · · · b(sk ). (4.8)
k! r(s1 ), · · · , r(sk ) α(t)
k=1 (t,S)
The second sum in (4.8) is over all partitions of t into k subtrees {s1 , · · · , sk }.
By (4.8), we can define relation between coefficients a(t) and b(t), See Table 4.1
to Table 4.4[LQ01] .
According to Table 4.1 – Table 4.4, we can determine modified equation for R–K
equation (up to 5 orders, it is clear that as long as the order continues to add 6, 7-order
tree · · · equation can be modified to any order).
Remark 4.4. If numerical method is symmetrical (or time-reversible), then when r(t)
is even, b(t) = 0.
Remark 4.5. If numerical method is p-order, in other words when r(t) ≤ p, a(t) =
1; then, if 2 ≤ r(t) ≤ p, b(•) = 1, b(t) = 0; if r(t) = p + 1, b(t) = a(t) − 1.
336 7. Symplectic Runge–Kutta Methods
•• • • • • • • • • • •
t= a =b + 2b( • )b + 2b( • )2 b( •/ ) + b( • )4
• • • •
• % • & % • & • • 2 •
4 •
t=• • a • • =b • • + b( • )b + b( • )b • + b( / )2
3 • 3 • •
• • •
10 •
b( • )2 b( •/ ) + b( • )4
+
3
• • %• •& %• •& • • •
t= • a • =b • + 2b( • ) b + 2b( • )b •
• •
• • •
•
+4b( • )2 b( •/ ) + b(•)4
• %• & %• & •
• • • •
t=• a =b + 4b( • ) b • + 3b( / )2
• • • •
• • •
•
+6b( • )2 b( •/ ) + b( • )4
z + z
n n+1
zn+1 = zn + hf .
2
" h2 (2) h4 7 (4)
z) −
ż = f (" f (f, f ) − 2f (1) f (1) f + f (f, f, f, f)
24 120 48
1 1 3
+ f (3) (f, f, f (1) f ) − f (2) f, f (2) (f, f ) − f (2) (f, f (1) f (1) f )
4 4 2
3 7 (1) (3) 1
+ f (2) (f (1) f, f (1) f ) − f f (f, f, f ) − f (1) f (2) (f, f (1) f )
4 12 2
1 3
+ f (1) f (1) f (2) (f, f ) + f (1) f (1) f (1) f (1) f + O(h6 ).
4 2
1
b(t) by Section 7.5.
7.4 Formal Energy for Symplectic R–K Method 337
5 ••• • •
• • • • • • 10
t = • • a • • = b • • + b( • )b( )+ b( • )2 b( )
• • • 2 • 3 •
5 •
+ b ( • )3 b( •/ ) + b( • )5
2
• • • % • • •& % • • •& • • 5 • • • 10 •
5
t= • a • =b • + b( • )b • + b( • )b + b( • )2 b •
• • • 2 • 2 • 3 •
10 • • •
+ b( • )2 b + 5b( • )3 b( /
• ) + b( • )
5
3 •
• •
+15b( • )b( •/ )2 + 10b( • )3 b( •/ ) + b( • )5
√ √
1 3 1 1 3
− −
2 6 4 4 6
√ √
1 3 1 3 1
+ +
2 6 4 6 4
1 1
2 2
Modified equation is defined2
1 1
0
4 4
3 1 1
4 2 4
1 1
2 2
s
Lemma 4.9. Let (A, b) be a symplectic R–K method, a(t) = γ(t) bj φj (t), then:
j=1
ż = f"("
" " z̃ .
z) = J H (4.11)
Conversely, if modified equation of symplectic R–K the method (A, b) are f"("
z ), then
we have
H" z = −J f"(z). (4.12)
Proof. Note that both the modified equation and the formal energy can be obtained
essentially via series expansion, Lemma 4.10 is obvious.
s
Lemma 4.11. Let (A, b) be a symplectic R–K method, a(t) = γ(t) bj φj (t), then:
j=1
b(u ◦ v) b(v ◦ u)
+ = 0, u, v ∈ T P, u = v, (4.13)
γ(u ◦ v) γ(v ◦ u)
Proof. Using the method that proves Lemma 10 in literature[Hai94] and modifying it
slightly will complete the proof. We leave out its detail here.
We now need another coefficient ν(t) related to rooted tree. Note first that the
rooted trees u ◦ v and v ◦ u represent an identical unrooted tree with different root, i.e.,
selecting the other vertex of u ◦ v as root leads to tree v ◦ u. Thus, in an equivalent
class, it can always transform one rooted tree to another by selecting a different root.
Take t̄ ∈ T E, then t̄ is one equivalent class of rooted tree collection (t represents the
element). Let u ∈ t̄, then u can be obtained by selecting some vertex of t as the root
node. We denote ν(u) as number of vertices of t which may be selected.
Example 4.13. t = [τ, τ, [τ ]], then t̄ = {t1 , t2 , t3 , t4 }, where •
• • • • • • •
t1 = • • • , t2 = • • , t3 = • , t4 = • .
• • • •
We have •
7.4 Formal Energy for Symplectic R–K Method 341
F ∗ (u ◦ v) = −F ∗ (v ◦ u). (4.16)
Proof. First the right side of (4.17) is uniquely determined, and on the left side the
selection of t∗ may not necessarily be unique. Therefore, it is required to prove that
(4.17) the left side of the formula is independent of selection of t∗ . We explain it as
follows: given t∗1 ∈ t∗ , such that σ(t∗ ) = σ(t∗1 ), then there exists a series of u ◦ v,
such that
F ∗ (u ◦ v) = −F ∗ (v ◦ u) =⇒ ∇F ∗ (u ◦ v) = −∇F ∗ (v ◦ u),
therefore
α(t∗ )b(t∗ )∇F ∗ (t∗ ) = α(t∗1 )b(t∗1 )∇F ∗ (t∗1 ). (4.19)
∗1
For m > 2, t must be a node of (4.18), (4.19) is also held.
Next, ∇F ∗ (t∗ ) must be a linear combination of basic differentials in the same
class, i.e.,
342 7. Symplectic Runge–Kutta Methods
∇F ∗ (t∗ ) = β(t)F (t).
t∈t¯∗
l m
t∗ = • • •j ,
•i
Thus, we get:
It is easy to see,
ν(t)
β(t) = ± ,
ν(t∗ )
where “ ± ” is selected using the following rule: if d(·) expresses the distance between
the vertex to the root node, i.e., the least number of vertices passed from this vertex to
the root node along the connection between vertices (including initial point and root
node), then sign (β(t)) = (−1)d(t) . Using the above example
sign (b(t))
(−1)d(t) = ,
sign (b(t∗ ))
therefore:
α(t)b(t)
∇F ∗ (t∗ ) = F (t).
¯∗
α(t∗ )b(t∗ )
t∈t
7.4 Formal Energy for Symplectic R–K Method 343
Thus, we get
α(t∗ )b(t∗ )∇F ∗ (t∗ ) = α(t)b(t)F (t).
t∈t¯∗
With the above results, we describe the main result of this section[Hai94] .
Theorem 4.16. Given a R–K method (A, b), A = (aij )s×s , b = (b1 , b2 , · · · , bs ) , its
formal energy is
hρ(t)−1
" h) = −J
H(z, α(t)b(t)F ∗ (t), "
t ∈ T E, (4.20)
ρ(t)!
ρ(t)≤N
where b(t) is determined by a(t) (According to Table 4.1 to Table 4.4), i.e.,
s
a(t) = γ(t) bj φj (t).
j=1
then
hr(t∗ )−1
"
ż = α(t)b(t)F (t)("
z)
r(t∗ ) ! ¯∗
"
t∗ ∈T E t∈t
hr(t∗ )−1
= α(t∗ )b(t∗ )∇F ∗ (t∗ ).
r(t∗ ) !
"
t∗ ∈T E
By Lemma 4.10
hr(t∗ )−1
" z = −J
H α(t∗ )b(t∗ )∇F ∗ (t∗ ),
r(t∗ )!
"
t∗ ∈T E
Remark 4.17. Literature [Tan94] pointed out that each item of series expansion of for-
mal energy of symplectic R–K scheme has 1 to 1 corresponding relationship with
unrooted trees collection. This theorem specifically indicates this 1 to 1 correspon-
dences.
344 7. Symplectic Runge–Kutta Methods
Finally we sum up the method to construct the formal energy of a symplectic R–K
method: given a symplectic R–K method (A, b, c), let a(t) = γ(t) bj φj (t). Then
j
according to Table 4.1 to Table 4.4 identify the corresponding b(t) to each rooted
" Using (4.20), we can directly write the formal energy. Without loss of
tree in T E.
generality, in practice we can choose
7
• • • • • •
"= •, • • • • • • • , • • , ··· .
TE , , • •,
• • • • •
If we know the order of method or the method are time reversible (symmetrical),
then according to Remark 4.4 and Remark 4.5 many calculations can be left out.
1 1
0
4 4
3 1 1
4 2 4
1 1
2 2
7.5 Definition of a(t) and b(t) 345
1
2
• •
a( • ) = 1 a(•/ ) = 1 b( • ) = 1 b •/ = 0
3 • 3 • 1
a • • = b • • =−
1
a • = b • =
• 4 • 2 • 4 • 2
• • • 1 • • • • •
a = a • • =1 b =0 b • • =0
• 2 •
• •
• • • • • •
a • • =3 b • • =0
3
a • = b • =0
• 2 • • •
• •
a •• •• = b •• •• =
5 5 7 1
a ••• = b ••• =
• 16 • 8 • 48 • 24
• •
• • • • • •
a •• = b •• =−
15 15 1 3
a • • = b • • =−
16 8 16 8
• • • •
346 7. Symplectic Runge–Kutta Methods
• • 5 • • • • • 1 • • •
5 7
a • • = a • = b • • = b • =−
•
4 • 4
•
4 • 12
• •
• • • • • • • •
5 15 1 1
a • = a • = b • =− b • =
• 2 • 4 • 6 • 4
• •
• • • •
15 3
a • = b • =
• 2 • 2
• •
√ √
1 3 1 1 3
− −
2 6 4 4 6
√ √
1 3 1 3 1
+ +
2 6 4 6 4
1 1
2 2
a( • ) = 1 a( /•) = 1 b( • ) = 1 b /• = 0
• •
• •
a( • • ) = 1 a( •) = 1 b • • =0 b • =0
• • • •
• • • • • • •
••• • •
a( )=1 a =1 b =0 b =0
• • • •
• • • •
a • • =1 a • • =1 b • • =0 b • • =0
• • • •
• • 35 • • • •
35 37 1
a • • = 72 a ••• = b • • = − 72 b ••• =−
• • 36 • • 36
• • 35 • • 37
a • • = b • • =−
72 72
• •
7.6 Multistep Symplectic Method 347
1 1
0
4 4
3 1 1
4 2 4
1 1
2 2
a( • ) = 1 a( /•) = 1 b( • ) = 1 b /• = 0
• •
• • 1
a( • • ) = b • • =−
15 9 1
a •) = b • =
• 16 • 8 • 16 • 8
• • • • • • •
••• 1
a( )= a • =1 b =0 b • =0
• 2 • • •
• • • •
a • • =3 b • • =0
3
a • • = b • • =0
•
2 • • •
• • • • • •
205 115 51 13
a • • = 256 a ••• = b • • = − 256 b ••• =−
• • 128 • • 128
• • • •
65 1
a • • = b • • =−
64 64
• •
where z = (z1 , · · · , zn ) and a(z) = (a1 (z), · · · , an (z)) is a smooth vector field on
Rn . For Equations (6.1) we define a linear m step method (LMM) in standard form
by
m m
αj zj = τ βj Qj , (6.2)
j=0 j=0
αm = 1, |α0 | + |β0 | = 0.
348 7. Symplectic Runge–Kutta Methods
αm zm + · · · + α1 z1 + α0 z0 = τ a(βm zm + · · · + β1 z1 + β0 z0 ). (6.6)
Our goal is to find a matrix g, i.e., a linear transformation g : R2n → R2n which
can satisfy (6.6)
αm g m (z0 ) + · · · + α1 g(z0 ) + α0 z0 = τ a βm g m (z0 ) + · · · + β1 g(z0 ) + β0 z0 . (6.7)
Such a map g exists for sufficiently small τ and can be represented by continued
fractions and rational approximations. We call this transformation step transition
operator[Fen98b] .
Definition 6.1. If g is a symplectic transformation, then its corresponding LMM (6.6)
is symplectic (we simply call the method SLMM).
7.6 Multistep Symplectic Method 349
α0 I + α1 g 1 + · · · + αm g m
τa = . (6.8)
β0 I + β1 g 1 + · · · + βm g m
where μ is the eigenvalue of the infinitesimal symplectic matrix a and λ is the eigen-
value of g.
Let
ξ(λ)
ψ(λ) = , (6.10)
σ(λ)
then (6.9) can be written as
τ μ = ψ(λ). (6.11)
Its inverse function is
λ = φ(τ μ). (6.12)
To study the symplecticity of the LMM, one only needs to study the properties
of functions φ and ψ. We will see that if φ is of the exponential form or ψ is of
logarithmic form, the corresponding LMM is symplectic. We first study the properties
of the exponential functions and logarithmic functions.
Explike and loglike functions
First we describe the properties of exponential functions:
(1) exp (x)|x=0 = 1.
d
(2) exp (x)|x=0 = 1.
dx
(3) exp (x + y) = exp (x) · exp (y).
If we substitute y with −x, we have
m
m
1
we obtain ψ(λ) + ψ = 0. Now ξ(1) = αk = 0, σ(1) = βk = 0, then
λ
k=0 k=0
ξ(1)
ψ(1) = = 0.
σ(1)
7.6 Multistep Symplectic Method 351
Corollary 6.7. If above generating polynomial is consistent with ODE (6.1), then
1
ψ(λ) is loglike function, i.e., ψ + ψ(λ) = 0, ψ(1) = 0, ψ̇(1) = 1.
λ
ξ̇σ − ξ̇ξ ξ̇(1)
Proof. ψ (1) = = = 1. This condition is just consistence condition.
σ2 σ(1)
ξ(λ)
Theorem 6.8. Let ψ(λ) = irreducible loglike function, then ξ(λ) is an anti-
σ(λ)
symmetric polynomial while σ(λ) is a symmetric one.
Proof. We write formally
If deg ξ(λ) = p < m, set ai = 0 for i > p; if deg σ(λ) = q < m, set βi = 0 for
ξ(1)
i > q. ψ(1) = 0 ⇒ ξ(1) = 0, since otherwise, if ξ(1) = 0, then ψ(1) = = 0.
σ(1)
Now ξ(1) = 0 ⇔ σ(1) = 0, since otherwise ξ(1) = σ(1) ⇒ ξ(λ), σ(λ) would have
common factor. So we have
m
p
ξ(1) = αk = αk = 0,
k=0 k=0
m q
σ(1) = βk = βk = 0.
k=0 k=0
ξ(λ) ξ(λ)
=− ⇐⇒ ξ(λ)σ(λ) = −ξ(λ)σ(λ)
σ(λ) σ(λ)
m
σ(λ) = −σ̄(λ), and βk = σ(1), then−σ̄(1) = σ(1) ⇔ σ(1) = 0, which leads to
k=0
a contradiction with the assumption σ(1) = 0. Therefore c = −1, i.e.,
˜
ξ(λ) = −ξ(λ), αj = −αm−j , j = 0, 1, · · · , m,
σ(λ) = σ̃(λ), βj = βm−j , j = 0, 1, · · · , m.
The proof for the case m = deg σ(λ) proceeds in exactly the same manner as above.
Let z1 = cz0 , then z0 = c−1 z1 , insert this equation into (6.16),we get
% &
1 1 1 z2
z2 = 2τ az1 + z1 = 2τ a + z1 = d1 z1 , z1 = z = ,
c c 1 2 d1
2τ a +
c
⎛ ⎞
⎜ 1 ⎟ 1
z3 = z1 + 2τ az2 = ⎝2τ a +
1⎠ 2
z = d2 z2 , z2 = z3 ,
1
2τ a + 2τ a +
c 1
2τ a +
c
⎛ ⎞
⎜ ⎟
⎜ ⎟
⎜ 1 ⎟
z4 = ⎜2τ a + ⎟ = d4 z3 , ···.
⎜ 1 ⎟
⎝ 2τ a +
1 ⎠
2τ a +
c
2. Exponential function
∞
zk
exp(z) = 1 + . (6.19)
k!
k=1
We have
P0 p0 p1 1+z p2 P1 2+z
= = 1, = , = = ,
Q0 q0 q1 1 q2 Q1 2−z
(6.22)
p3 6 + 4z + z 2 p4 P2 12 + 6z + z 2
= , = = + ···.
q3 6 − 2z q4 Q2 12 − 6z + z 2
In general p2n−1 (z) is a polynomial of degree n, q2n−1 is a polynomial of degree n−1,
p2n−1
so is not explike. While p2n = Pn (x), q2n = Qn (x) are both polynomials of
q2n−1
degree n and from the recursions
is explike and
354 7. Symplectic Runge–Kutta Methods
and
a2n−1 = (n − 1)(w − 1), a2n = n(w − 1), n ≥ 2,
b2n−1 = 2n − 1, b2n = 2, n ≥ 2,
and the Euler’s contracted expansion
2 2
2(w − 1) 2(w − 1) 2 × 2(w − 1) 2(n − 1)(w − 1)
log w =
w + 1 – 6(w + 1) – 2.5(w + 1) – · · · – 2(2n − 1)(w + 1) – · · ·
A1 A2 A3 An
= , (6.27)
B1 + B2 + B3 + · · · + Bn + · · ·
where
A1 = 2(w − 1), A2 = −2(w − 1), · · · , An = −(2(n − 1)(w − 1))2 , n ≥ 3,
B0 = 0, B1 = w + 1, B2 = 6(w + 1), · · · , Bn = 2(2n − 1)(w + 1), n ≥ 2.
and for n ≥ 3,
Pn (w) = −(2(n − 1)(w − 1)2 Pn−2 (w) + 2(2n − 1)(w − 1)Pn−2 (w)),
(6.30)
Qn (w) = −((2n − 1)(w − 1)2 Qn−2 (w) + 2(2n − 1)(w − 1)Qn−2 (w)).
3(λ2 − 1)
So we see R1 (λ) is just the Euler midpoint rule and R2 (λ) = 2 is just the
λ + 4λ + 1
Simpson scheme.
Conclusion: The odd truncation of the continued fraction of the Lagrange’s ap-
proximation to exp(x) and log (x) is neither explike nor loglike, while the even trun-
cation is explike and loglike. The truncation of the continued fraction obtained from
Euler’s contracted expansion is explike and loglike.
4. Obreschkoff formula
Another rational approximation to a given function is the Obreschkoff formula[Obr40] :
n
Ckn
m
Ckm
Rm,n (x) = (x0 − x)k f (k) (x) − k
(x − x0 )k f (k) (x0 )
Ckm+n k! C k!
k=0 k=0 m+n
- x
1
= (x − t)m (x0 − t)n f m+n+1 (t)dt. (6.31)
(m + n)! x0
.
(1) Take f (x) = ex , x0 = 0, we obtain Padé approximation exp(x) = Rm,n (x).
If m = n, we obtain Padé approximation Rm,m (x).
.
(2) Take f (x) = log(x), x0 = 1, we obtain log(x) = Rm,n (x). If m = n, we
obtain loglike function Rm (x),
1 Ckm
m
Rm (λ) = k
(λ − 1)k (λm−k + (−1)k−1 λm ),
λm C k
k=1 2m
i.e., % &
1
Rm (λ) + Rm = 0.
λ
We have
356 7. Symplectic Runge–Kutta Methods
z = g 0 (z),
z1 = g(x),
z2 = g(g(z)) = g ◦ g(z) = g 2 (z), (6.32)
..
.
zn = g(g(· · · (g(z)) · · ·)) = g ◦ g ◦ · · · ◦ g ◦ (z) = g n (z),
It’s easy to prove that if LMM (6.33) is consistent with Equation (6.1), then for smooth
f and sufficiently small step-size τ , the operator g defined by (6.32) exists and it can
be represented as a power series in τ and with first term equal to identity. Consider a
case where Equation (6.1) is a Hamiltonian system, i.e., a(z) = J∇H(z), we have
the following definition.
This definition is a completely different criterion that can include the symplec-
tic condition for one-step methods in the usual sense. But Tang in[Tan93a] has proved
that nonlinear multistep method can satisfy such a strict criterion. Numerical experi-
ments of Li[Fen92b] show that the explicit 3-level centered method (Leap-frog method)
1
is symplectic for linear Hamiltonian systems H = (p2 + 4q 2 ) (see Fig. 0.2 in
2
introduction of this book) but is non-symplectic for nonlinear Hamiltonian systems
1 2
H = (p2 + q 2 ) + q 4 (see Fig. 0.3 (a,b,c) in introduction of this book).
2 3
Bibliography
[AS93] L. Abia and J.M. Sanz-Serna: Partitioned Runge–Kutta methods for separable Hamil-
tonian problems. Math. Comp., 60:617–634, (1993).
[But87] J.C. Butcher: The Numerical Analysis of Ordinary Differential Equations. John Wiley,
Chichester, (1987).
[CH95] M.P. Calvo and E. Hairer: Further reduction in the number of independent order condi-
tions for symplectic, explicit partitioned Runge-Kutta and Runge–Kutta–Nyström methods.
Appl. Numer. Math., 18:107–114, (1995).
[Chi97] S. A. Chin: Symplectic integrators from composite operator factorization. Physics
Letters A, 226:344–348, (1997).
[Coo87] G. J. Cooper: Stability of Runge–Kutta methods for trajectory problems. IMA J.
Numer. Anal., 7:1–13, (1987).
[CS93] M.P. Calvo and J.M. Sanz-Serna: High-order symplectic Runge-Kutta-Nyström meth-
ods. SIAM J. Sci. Comput., 114:1237–1252, (1993).
[CS94] M.P. Calvo and J.M. Sanz-Serna: Canonical B-Series. Numer. Math., 67:161–175,
(1994).
[DV84] K. Dekker and J.G. Verwer: Stability of Runge-Kutta Methods for Stiff Initial Value
Problems. North-Holland, Amsterdam, (1984).
[Fen65] K. Feng: Difference schemes based on variational principle. J. of Appl. and Comput.
Math.in Chinese, 2(4):238–262, (1965).
[Fen85] K. Feng: On difference schemes and symplectic geometry. In K. Feng, editor, Pro-
ceedings of the 1984 Beijing Symposium on Differential Geometry and Differential Equa-
tions, pages 42–58. Science Press, Beijing, (1985).
[Fen86a] K. Feng: Canonical Difference Schemes for Hamiltonian Canonical Differential
Equations. In International Workshop on Applied Differential Equations (Beijing, 1985),
pages 59–73. World Sci. Publishing, Singapore, (1986).
[Fen86b] K. Feng: Difference schemes for Hamiltonian formalism and symplectic geometry.
J. Comput. Math., 4:279–289, (1986).
[Fen86c] K. Feng: Symplectic geometry and numerical methods in fluid dynamics. In F.G.
Zhuang and Y.L. Zhu, editors, Tenth International Conference on Numerical Methods in
Fluid Dynamics, Lecture Notes in Physics, pages 1–7. Springer, Berlin, (1986).
[Fen91] K. Feng: The Hamiltonian Way for Computing Hamiltonian Dynamics. In R. Spigler,
editor, Applied and Industrial Mathematics, pages 17–35. Kluwer, The Netherlands, (1991).
[Fen92a] K. Feng: Formal power series and numerical methods for differential equations. In
T. Chan and Z.C. Shi, editors, International conf. on scientific computation, pages 28–35.
World Scientific, Singapore, (1992).
[Fen92b] K. Feng: How to compute property Newton’s equation of motion. In L. A. Ying,
B.Y. Guo, and I. Gladwell, editors, Proc of 2nd Conf. on Numerical Method for PDE’s,
pages 15–22.World Scientific, Singapore, (1992). Also see Collected Works of Feng Kang.
Volume I, II. National Defence Industry Press, Beijing, (1995).
[Fen93a] K. Feng: Formal dynamical systems and numerical algorithms. In K. Feng and
Z.C Shi, editors, International conf. on computation of differential equationsand dynamical
systems, pages 1–10. World Scientific, Singapore, (1993).
358 Bibliography
[Fen93b] K. Feng: Symplectic, contact and volume preserving algorithms. In Z.C. Shi and
T. Ushijima, editors, Proc.1st China-Japan conf. on computation of differential equation-
sand dynamical systems, pages 1–28. World Scientific, Singapore, (1993).
[Fen95] K. Feng: Collected Works of Feng Kang. volume I,II. National Defence Industry
Press, Beijing, (1995).
[Fen98a] K. Feng: The calculus of generating functions and the formal energy for Hamiltonian
systems. J. Comput. Math., 16:481–498, (1998).
[Fen98b] K. Feng: The step-transition operator for multi-step methods of ODEs. J. Comput.
Math., 16(3), (1998).
[FQ87] K. Feng and M.Z. Qin: The symplectic methods for the computation of Hamiltonian
equations. In Y. L. Zhu and B. Y. Guo, editors, Numerical Methods for Partial Differential
Equations, Lecture Notes in Mathematics 1297, pages 1–37. Springer, Berlin, (1987).
[FQ91a] K. Feng and M.Z. Qin: Hamiltonian Algorithms for Hamiltonian Dynamical Systems.
Progr. Natur. Sci., 1(2):105–116, (1991).
[FQ91b] K. Feng and M.Z. Qin: Hamiltonian algorithms for Hamiltonian systems and a com-
parative numerical study. Comput. Phys. Comm., 65:173–187, (1991).
[FQ03] K. Feng and M. Q. Qin: Symplectic Algorithms for Hamiltonian Systems. Zhejiang
Science and Technology Publishing House, Hangzhou, in Chinese, First edition, (2003).
[FW91a] K. Feng and D.L. Wang: A note on conservation laws of symplectic difference
schemes for Hamiltonian systems. J. Comput. Math., 9(3):229–237, (1991).
[FW91b] K. Feng and D.L. Wang: Symplectic difference schemes for Hamiltonian systems in
general symplectic structure. J. Comput. Math., 9(1):86–96, (1991).
[FW94] K. Feng and D.L. Wang: Dynamical systems and geometric construction of algo-
rithms. In Z. C. Shi and C. C. Yang, editors, Computational Mathematics in China, Con-
temporary Mathematics of AMS Vol 163, pages 1–32. AMS, (1994).
[FW98] K. Feng and D.L. Wang: On variation of schemes by Euler. J. Comput. Math., 16:97–
106, (1998).
[FWQ90] K. Feng, H.M. Wu, and M.Z. Qin: Symplectic difference schemes for linear Hamil-
tonian canonical systems. J. Comput. Math., 8(4):371–380, (1990).
[FWQW89] K. Feng, H.M. Wu, M.Z. Qin and D.L. Wang: Construction of canonical dif-
ference schemes for Hamiltonian formalism via generating functions. J. Comput. Math.,
7:71–96, (1989).
[Ge88] Z. Ge: Symplectic geometry and its application in numerical analysis. PhD thesis,
Computer Center, CAS, (1988).
[Ge90] Z. Ge: Generating functions, Hamilton–Jacobi equations and symplectic groupoids on
Poisson manifolds. Indiana Univ. Math. J., 39:859, (1990).
[Ge91] Z. Ge: Equivariant symplectic difference schemes and generating functions. Physica
D, 49:376–386, (1991).
[Ge95] Z. Ge: Symplectic integrators for Hamiltonian systems. In W. Cai et al., editor, Nu-
merical Methods in Applied Sciences, pages 97–108, Science Press, New York, (1995).
[Gon96] O. Gonzalez: Time integration and discrete Hamiltonian systems. J. Nonlinear. Sci.,
6:449–467, (1996).
[Hai94] E. Hairer: Backward analysis of numerical integrators and symplectic methods. Annals
of Numer. Math., 1:107–132, (1994).
[Hai97b] E. Hairer: Variable time step integration with symplectic methods. Appl. Numer.
Math., 25:219–227, (1997).
[Hai99] E. Hairer: Backward error analysis for multistep methods. Numer. Math., 84:199–232,
(1999).
[Hai00] E. Hairer: Symmetric projection methods for differential equations on manifolds. BIT,
40:726–734, (2000).
[Hai01] E. Hairer: Geometric integration of ordinary differential equations on manifolds. BIT,
41:996–1007, (2001).
[Hai03] E. Hairer: Global modified Hamiltonian for constrained symplectic integrators. Nu-
mer. Math., 95:325–336, (2003).
Bibliography 359
[Hen62] P. Henrici: Discrete Variable Methods in Ordinary Differential Equations. John Wiley
& Sons, Inc., New York, Second edition, (1962).
[HL97a] E. Hairer and P. Leone: Order barriers for symplectic multi-value methods. In D.F.
Grifysis, D.F.Higham, and G.A. Watson, editors, Numerical Analysis 1997 Proc.of the 17-
th Dundee Biennial Conference,June 24-27, 1997, Pitman Reserch Notes in math. series
380, pages 133–149, (1997).
[HL97b] E. Hairer and Ch. Lubich: The life-span of backward error analysis for numerical
integrators. Numer. Math., 76:441–462, (1997).
[HL97c] M. Hochbruck and Ch. Lubich: On Krylov subspace approximations to the matrix
exponential operator. SIAM J. Numer. Anal., 34(5), (1997).
[HL97d] W. Huang and B. Leimkuhler: The adaptive Verlet method. SIAM J. Sci. Comput.,
18(1):239, (1997).
[HL99a] E. Hairer and Ch. Lubich: Invariant tori of dissipatively perturbed Hamiltonian sys-
tems under symplectic discretization. Appl. Numer. Math., 29:57–71, (1999).
[HL99b] M. Hochbruck and Ch. Lubich: Exponential integrators for quantum-classical molec-
ular dynamics. BIT, 39:620–645, (1999).
[HL00a] E. Hairer and P. Leone: Some properties of symplectic Runge–Kutta methods. New
Zealand J. of Math., 29:169–175, (2000).
[HL00b] E. Hairer and Ch. Lubich: Energy conservation by Störmer-type numerical inte-
grators. In G.F. Griffiths and G.A. Watson, editors, In Numerical Analysis 1999, pages
169–190. CRC Press LLC, (2000).
[HL00c] E. Hairer and Ch. Lubich: Long-time energy conservation of numerical methods for
oscillatory differential equations. SIAM J. Numer. Anal., 38:414–441, (2000).
[HL00d] J. L. Hong and Y. Liu: Symplectic integration of linear discontinues Hamiltonian
systems. Neural Parallel Sci Comput., 8:317–325, (2000).
[HL03] M. Hochbruck and C. Lubich: On magnus integrators for time-dependent Schrödinger
equations. SIAM J. Numer. Anal., 41:945–963, (2003).
[HL04a] E. Hairer and C. Lubich: symmetric multistep methods over long times. Numer.
Math., 97:699–723, (2004).
[HLR01] T. Holder, B. Leimkuhler, and S. Reich: Explicit variable step-size and time-
reversible integration. Appl. Numer. Math., 39:367–377, (2001).
[HLS98] M. Hochbruck, C. Lubich, and H. Selhofer: Exponential integrators for large systems
of differential equations. SIAM J. Sci. Comput., 19(5):1552–1574, (1998).
[HLW02] E. Hairer, Ch. Lubich, and G. Wanner: Geometric Numerical Integration. Num-
ber 31 in Springer Series in Computational Mathematics. Springer-Verlag, (2002).
[HLW03] E. Hairer, C. Lubich and G. Wanner: Geometric integration illustrated by the
Störmer-Verlet method. Acta Numerica, pages 399–450, (2003).
[HM04] P. Hydon and E.L. Mansfield: A variational complex for difference equations. Foun-
dations of Computational Mathematics, 4:187–217, (2004).
[HMM95] P. Hut, J. Makino and S. McMillan: Building a better leapfrog. Astrophys. J.,
443:L93–L96, (1995).
[HMSS93] E. Hairer, A. Murua and J.M. Sanz-Serna: The non-existence of symplectic multi-
derivative Runge–Kutta methods. Preprint, (1993).
[HNW93] E. Hairer, S. P. Nørsett, and G. Wanner: Solving Ordinary Differential Equations I,
Nonstiff Problems. Springer-Verlag, Second revised edition, (1993).
[HOS99] D.J. Hardy, D.I. Okunbor, and R.D. Skeel: Symplectic variable step size integration
for n-body problems. Appl. Numer. Math., 29:19–30, (1999).
[HS81] W. H. Hundsdorfer and M. N. Spijker: A note on B-stability of Runge–Kutta methods.
Numer. Math., 36:319–331, (1981).
[HS94] A. R. Humphries and A. M. Stuart: Runge-Kutta methods for dissipative and gradient
dynamical systems. SIAM J. Numer. Anal., 31(5):1452–1485, (1994).
[HS97a] E. Hairer and D. Stoffer: Reversible long-term integration with variable stepsizes.
SIAM J. Sci. Comput., 18:257–269, (1997).
360 Bibliography
[HS05] E. Hairer and G. Söderlind: Explicit time reversible adaptive step size control. SIAM
J. Sci. Comput., 26:1838–1851, (2005).
[HW74] E. Hairer and G. Wanner: On the Butcher group and general multivalue methods.
Computing, 13:1–15, (1974).
[HW81] E. Hairer and G. Wanner: Algebraically stable and implementable Runge–Kutta meth-
ods of high order. SIAM J. Numer. Anal., 18:1098–1108, (1981).
[HW91] E. Hairer and G. Wanner: Solving Ordinary Differential Equations II, Stiff and
Differential-Algebraic Problems. Springer, Berlin, (1991).
[HW94] E. Hairer and G. Wanner: Symplectic Runge-Kutta methods with real eigenvalues.
BIT, 34:310–312, (1994).
[HW96] E. Hairer and G. Wanner: Solving Ordinary Differential Equations II. Stiff and
Differential-Algebraic Problems, 2nd edition, Springer Series in Computational Mathemat-
ics 14. Springer-Verlag Berlin, Second edition, (1996).
[IA88] T. Itoh and K. Abe: Hamiltonian-conserving discrete canonical equations based on
variational difference quotients. J. of Comp. Phys., 76:85–102, (1988).
[Jay96] L. O. Jay: Symplectic partitioned Runge–Kutta methods for constrained Hamiltonian
systems. SIAM J. Numer. Anal., 33:368–387, (1996).
[Jay97] L. O. Jay: Lagrangian integration with symplectic methods. Technical Report AH-
PCRC Preprint 97-009, University of Minnesota, (1997).
[Jay99] L. O. Jay: Structure preservation for constrained dynamics with super partitioned
additive Runge–Kutta methods. SIAM J. Sci. Comput., 20(2):416–446, (1999).
[Jim94] S. Jiménez: Derivation of the discrete conservation laws for a family of finite differ-
ence schemes. Applied Mathematics and Computation, 64:13–45, (1994).
[JL06] Z. Jia and B. Leimkuhler: Geometric integrators for multiple time-scale simulation. J.
Phys. A: Math. Gen., 39:5379–5403, (2006).
[Kar96a] B. Karasözen: Comparison of reversible integrators for a Hamiltonian in normal
form. In E. Kreuzer and O. Mahrenholz, editors, Proceedings of the Third International
Congress on Industrial and Applied Mathematics, ICIAM 95, Issue 4: Applied Sciences,
especially Mechanics (Minisymposia), pages 563–566, (1996).
[Kar96b] B. Karasözen: Composite integrators for Bi-Hamiltonian systems. Comp. & Math.
with Applic., 32:79–86, (1996).
[Kar96c] B. Karasözen: Numerical Studies on a Bi-Hamiltonian Hénon-Heiles System. Tech-
nical Report No 133, Middle East Technical University, Department of Mathematics,
Ankara, Turkey, (1996).
[Kar97] B. Karasözen: Reflexive methods for dynamical systems with conserved quantities.
Technical Report Nr. 1897, Technische Hochschule Darmstadt, FB Mathematik, (1997).
[KHL08] L. H. Kong, J. L. Hong, and R. X. Liu: Long-term numerical simulation of the
interaction between a neutron field and meson field by a symplectic-preserving scheme. J.
Phys. A: Math. Theor., 41:255207, (2008).
[Kir86] U. Kirchgraber: Multi-step methods are essentially one-step methods. Numer. Math.,
48:85–90, (1986).
[Lam91] J.D. Lambert: Numerical Methods for Ordinary Differential Equations, The Initial
Value Problem. Wiley, Chichester, (1991).
[Las88] F.M. Lasagni: Canonical Runge–Kutta methods. Z. Angew. Math. Phys., 39:952–953,
(1988).
[LDJW00] Y.X. Li, P. Z. Ding, M. X. Jin, and C. X. Wu: Computing classical trajectories of
model molecule A2 B by symplectic algorithm. Chemical Journal of Chinese Universities,
15(8):1181–1186, (2000).
[Lei99] B. J. Leimkuhler: Reversible adaptive regularization: Perturbed Kepler motion and
classical atomic trajectories. Phil. Trans. Royal Soc. A, 357:1101, (1999).
[Leo00] P. Leone: Symplecticity and Symmetry of General Integration Methods. Thèse, Section
de Mathématiques, Université de Genève, Second edition, (2000).
[LP96] B. J. Leimkuhler and G. W. Patrick: A symplectic integrator for Riemannian manifolds.
J. Nonlinear. Sci., 6(4):367–384, (1996).
Bibliography 361
[LP01] L. Lopez and T. Politi: Applications of the cayley approach in the numerical solution
of matrix differential systems on quadratic groups. Appl. Numer. Math., 36:35–55, (2001).
[LQ01] H. W. Li and M. Z. Qin: On the formal energy of symplectic R–K method. Math.
Num. Sinica, 23:75–92, (2001).
[LQHD07] X.S. Liu, Y.Y. Qi, J. F. He, and P. Z. Ding: Recent progress in symplectic algorithms
for use in quantum systems. Communications in Computational Physics, 2(1):1–53, (2007).
[LR94a] B. Leimkuhler and S. Reich: Symplectic integration of constrained Hamiltonian sys-
tems. Math. Comp., 63:589–605, (1994).
[LR05] B. Leimkuhler and S. Reich: Simulating Hamiltonian Dynamics. Cambridge Univer-
sity Press, Cambridge, First edition, (2005).
[LvV97] B. J. Leimkuhler and E. S. van Vleck: Orthosymplectic integration of linear Hamil-
tonian systems. Numer. Math., 77:269–282, (1997).
[LW76] J. D. Lambert and I. A. Watson: Symmetric multistep methods for periodic initial
value problems. J. Inst. Maths. Applics., 18:189–202, (1976).
[LYC06] H. Liu, J.H. Yuan, J.B. Chen, H. Shou, and Y.M. Li: Theory of large-step depth
extrapolation. Chinese journal Geophys., 49(6):1779–1793, (2006).
[McL95c] R. I. McLachlan: On the numerical integration of ODE’s by symmetric composition
methods. SIAM J. Numer. Anal., 16:151–168, (1995).
[McL95d] R. I. McLachlan: On the numerical integration of ordinary differential equations by
symmetric composition methods. SIAM J. Sci. Comput., 16:151–168, (1995).
[McL96] R. I. McLachlan: More on Symplectic Correctors. In Jerrold E. Marsden, George
W. Patrick, and William F. Shadwick, editors, Integration Algorithms and Classical Me-
chanics, volume 10 of Fields Institute Communications. Fields Institute, American Mathe-
matical Society, July (1996).
[McL02] R. McLachlan: Splitting methods. Acta Numerica, 11:341–434, (2002).
[Mie89] S. Miesbach: Symplektische Phasenfluß approximation zur Numerischen Integration
Kanonischer Differentialgleichungen. Master’s thesis, Technische Universität München,
(1989).
[MP92] S. Miesbach and H.J. Pesch: Symplectic phase flow approximation for the numerical
integration of canonical systems. Numer. Math., 61:501–521, (1992).
[MPQ04] R.I. McLachlan, M. Perlmutter, and G.R.W. Quispel: On the nonlinear stability of
symplectic integrators. BIT, 44:99–117, (2004).
[MQ98a] R. I. McLachlan and G. R. W. Quispel: Generating functions for dynamical systems
with symmetries, integrals, and differential invariants. Physica D, 112:298–309, (1998).
[MQ98b] R.I. McLachlan and G.R.W. Quispel: Numerical integrators that preserve symme-
tries and reversing symmetries. SIAM J. Numer. Anal., 35:586–599, (1998).
[MQ02] R. I. McLachlan and G. R. W. Quispel: Splitting methods. Acta Numerica, 11:341–
434, (2002).
[MQ03] R.I. McLachlan and G.R.W. Quispel: Geometric integration of conservative polyno-
mial ODEs. Appl. Numer. Math., 45:411–418, (2003).
[MQ04] D.I. McLaren and G.R.W. Quispel: Integral-preserving integrators. J. Phys. A: Math.
Gen., 37:L489–L495, (2004).
[MQR98] R. I. McLachlan, G. R. W. Quispel, and N. Robidoux: A unified approach to Hamil-
tonian systems, Poisson systems, gradient systems, and systems with Lyapunov functions
and/or first integrals. Physical Review Letters, 81:2399–2403, (1998).
[MQR99] R. I. McLachlan, G. R. W. Quispel, and N. Robidoux: Geometric integration using
discrete gradients. Phil. Trans. Royal Soc. A, 357:1021–1046, (1999).
[MQT98] R. I. McLachlan, G. R. W. Quispel, and G. S. Turner: Numerical integrators that
preserve symmetries and reversing symmetries. SIAM J. Numer. Anal., 35(2):586–599,
(1998).
[MS95c] R. I. McLachlan and C. Scovel: Equivariant constrained symplectic integration. J.
Nonlinear. Sci., 5:233–256, (1995).
362 Bibliography
[Tan94] Y. F. Tang: Formal energy of a symplectic scheme for Hamiltonian systems and its
applications. Computers Math. Applic., 27:31–39, (1994).
[Vog56] R. de Vogelaere: Methods of integration which preserve the contact transformation
property of the Hamiltonian equations. Report No. 4, Dept. Math., Univ. of Notre Dame,
Notre Dame, Ind., Second edition, (1956).
[Wan91a] D. L. Wang: Semi-discrete Fourier spectral approximations of infinite dimensional
Hamiltonian systems and conservations laws. Computers Math. Applic., 21:63–75, (1991).
[Wan91b] D. L. Wang: Symplectic difference schemes for Hamiltonian systems on Poisson
manifolds. J. Comput. Math., 9(2):115–124, (1991).
[Wan91c] D. L. Wang: Poisson difference schemes for Hamiltonian systems on Poisson man-
ifolds. J. Comput. Math., 9:115–124, (1991).
[Wan93] D. L. Wang: Decomposition vector fields and composition of algorithms. In Proceed-
ings of International Conference on computation of differential equations and dynamical
systems, Beijing, 1993. World Scientific, (1993).
[Wan94] D. L. Wang: Some acpects of Hamiltonian systems and symplectic defference meth-
ods. Physica D, 73:1–16, (1994).
[Yos90] H. Yoshida: Conserved quantities of symplectic integrators for Hamiltonian systems.
Preprint, (1990).
[ZQ93a] M. Q. Zhang and M. Z. Qin: Explicit symplectic schemes to solve vortex systems.
Comp. & Math. with Applic., 26(5):51, (1993).
[ZQ93b] W. Zhu and M. Qin: Applicatin of higer order self-adjoint schemes of PDE’s. Com-
puters Math. Applic.,26(3):15–26, (1993).
[ZQ93c] W. Zhu and M. Qin: Constructing higer order schemes by formal power series. Com-
puters Math. Applic.,25(12):31–38, (1993).
[ZQ93] W. Zhu and M. Qin: Order conditionof two kinds of canonical difference schemes.
Computers Math. Applic., 25(6):61–74, (1993).
[ZQ94] W. Zhu and M. Qin: Poisson schemes for Hamiltonian systems on Poisson manifolds.
Computers Math. Applic., 27:7–16, (1994).
[ZQ95a] W. Zhu and M. Qin: Reply to “comment on Poisson schemes for Hamiltonian systems
on Poisson manifolds”. Computers Math. Applic., 29(7):1, (1995).
[ZQ95b] W. Zhu and M. Qin: Simplified order conditions of some canonical difference
schemes. J. Comput. Math., 13(1):1–19, (1995).
[ZS75] K. Zare and V. Szebehely: Time transformations in the extended phase-space. Celest.
Mech., 11:469–482, (1975).
[ZS95] M. Q. Zhang and R. D. Skeel: Symplectic integrators and the conservation of angular
momentum. J. Comput. Chem., 16:365–369, (1995).
[ZS97] M. Q. Zhang and R. D. Skeel: Cheap implicit symplectic integrators. Appl. Numer.
Math., 25(2):297, (1997).
[ZW99] H. P. Zhu and J. K. Wu: Generalized canonical transformations and symplectic algo-
rithm of the autonomous Birkhoffian systems. Progr. Natur. Sci., 9:820–828, (1999).
[ZzT96] W. Zhu, X. zhao, and Y Tang: Numerical methods with a high order of accuracy
applied in the quantum system. J. Chem. Phys., 104(6):2275–2286, (1996).
Chapter 8.
Composition Scheme
Ż = f (Z). (1.2)
would be of order 4 (i.e., Z3 − Z(t + h) = O(h5 )), where Z0 = Z(t). Z(t + h) is the
exact solution at t + h and Z3 the numerical one when the parameters c1 , c2 , and c3
are chosen properly.
We will use the method of Taylor expansion to deal with the simple case when
there is only one ordinary differential equation (ODE). When we deal with the case of
systems of ODEs, the Taylor expansions become very complex, although they surely
can be applied and the same conclusion as in the former case can be derived. We
introduce another method[HNW93] , known as “trees and elementary differentials” to
deal with the latter case. In fact, the essence of the two methods is the same; they are
just two different ways of expression.
In this section, without specific statements, the values of all functions and their
derivatives are calculated at Z0 , and we consider only the terms up to o(h4 ) in the
following calculations, while the higher order terms of h are omitted,
h2 h3 2 h4 3
Z(t+h) = Z0 +hf + f f + (f f 2 +f f )+ (f f 3 +4f f f 2 +f f )+O(h5 ).
2! 3! 4!
(1.5)
Now, we turn to the Taylor expansion of the numerical solution. We can rewrite
(1.3) as
We use the same technique to expand the Taylor expansions of f (Z2 ), f (Z3 ).
Since
Inserting the Taylor expansion of f (Z1 ) into right side of (1.9), we get
f (Z1 ) = f (Z0 ) + c1 hf f (Z0 ) + f (Z0 ) + c1 hf f (Z0 ) + f (Z1 )
2
+ f2 ! (c1 h)2 f (Z0 ) + f (Z1 ) + f2 ! (c1 h)2 f (Z0 ) + f (Z0 )
2
3
+c1 hf f (Z0 ) + f (Z1 ) + f3 ! (c1 h)3 f (Z0 ) + f (Z0 ) + O(h4 )
= f (Z0 ) + c1 hf 2f (Z0 ) + c1 hf f (Z0 ) + f (Z0 ) + c1 2hf f (Z0 )
2
+(c1 h)2 f2 ! f (Z0 ) + f (Z0 ) + (c1 h)2 f2 ! 2f (Z0 ) + c1 hf f (Z0 )
2
3
+f (Z0 ) + (c1 h)3 f3 ! 2f (Z0 ) + O(h4 )
2
= f (Z0 ) + c1 h 2f f (Z0 ) + (c1 h)2 2f f (Z0 ) + 2f f 2 (Z0 )
3
+(c1 h)3 2f f (Z0 ) + 6f f f 2 (Z0 ) + 43 f f 3 (Z0 ) + O(h4 ).
(1.10)
Similarly, developing f (Z2 ) and f (Z3 ), since
Z2 − Z0 = c1 h f (Z1 ) + f (Z0 ) + c2 h f (Z2 ) + f (Z1 )
= c1 hf (Z0 ) + (c1 + c2 )hf (Z1 ) + c2 hf (Z2 ),
therefore
f
f (Z2 ) = f (Z0 ) + hf c1 f (Z0 ) + (c1 + c2 )f (Z1 ) + c2 f (Z2 ) + h2 c1 f (Z0 )
2!
2 f
+(c1 + c2 )f (Z1 ) + c2 f (Z2 ) + h3 c1 f (Z0 ) + (c1 + c2 )f (Z1 )
3!
3
+c2 f (Z2 ) + O(h4 )
%
= f (Z0 ) + hf c1 f (Z0 ) + (c1 + c2 )f (Z1 ) + c2 f (Z0 ) + hf c1 f (Z0 )
+(c1 + c2 )f (Z1 ) + c2 f (Z0 ) + hf c1 f (Z0 ) + (c1 + c2 )f (Z1 ) + c2 f (Z0 )
f 2 & f %
+ h2 c1 f (Z0 ) + (c1 + c2 )f (Z1 ) + c2 f (Z0 ) + h2 c1 f (Z0 )
2! 2!
+(c1 + c2 )f (Z1 ) + c2 f (Z0 ) + hf c1 f (Z0 ) + (c1 + c2 )f (Z1 )
&2 f 3
+c2 f (Z2 ) + h3 c1 f (Z0 ) + (c1 + c2 )f (Z1 ) + c2 f (Z0 ) + O(h4 ).
3!
(1.11)
368 8. Composition Scheme
+(c1 + c2 + c3 )2 2f f 2 (Z0 ) + h3 c1 + c2 )c21 + (c2 + c3 )(c1 + c2 )2
3
+c3 (c1 + c2 + c3 )2 2f f (Z0 ) + (c1 + c2 )c21 + (c2 + c3 )(c1 + c2 )2
+c3 (c1 + c2 + c3 )2 + 2(c1 + c2 + c3 )3 2f f f 2 f (Z0 )
4
+ (c2 + c2 + c3 )3 f f 3 (Z0 ) + O(h4 ). (1.14)
3
w w
Let c1 = c3 = 1 , c2 = 0 , take into account (1.5) and (1.6), and compare the
2 2
Taylor expansion of the exact solution (1.5) with the above one. In order to get fourth
order accuracy schemes (1.3), we need to solve the following equations for coefficients
c1 , c2 , c3 :
1
+c3 (c1 + c2 + c3 )2 + 2(c1 + c2 + c3 )3 = . (1.20)
24
1 1
When 2w1 + w0 = 1 holds, the Equation (1.16) becomes = , i.e., identity, and
2 2
the Equations (1.17) – (1.20) become the same, i.e.,
Thus, we get the conditions for the difference scheme (1.3) to be of order 4:
2w1 + w0 = 1,
(1.21)
6w13 − 12w12 + 6w1 − 1 = 0.
Thus we get,
1
−2 3 1
w0 = 1 , w1 = 1 .
2 − 23 2 − 23
370 8. Composition Scheme
We use the “method of tree and elementary differentials” [HNW93] given in Chapter 7.
We first rewrite the scheme (1.3) in the R–K methods:
⎧
⎪
⎨ Z1 = Z0 + hc1 f (Z0 ) + c1 f (Z1 ) ,
Z2 = Z0 + h c1 f (Z0 ) + (c1 + c2 )f (Z1 ) + c2 f (Z2 ) ,
⎪
⎩
Z3 = Z0 + h c1 f (Z0 ) + (c1 + c2 )f (Z1 ) + (c2 + c3 )f (Z2 ) + c3 f (Z3 ) .
(1.23)
Obviously, the above equation is equivalent to following
⎧
⎪
⎪ g1 = Z0 ,
⎪
⎪
⎪
⎪ g
⎨ 2 = Z0 + c1 hf (g1 ) + c1 hf (g2 ),
g3 = Z0 + c1 hf (g1 ) + (c1 + c2 )hf (g2 ) + c2 hf (g3 ),
⎪
⎪
⎪
⎪ g4 = Z0 + c1 hf (g1 ) + (c1 + c2 )hf (g2 ) + (c2 + c3 )hf (g3 ) + c3 hf (g4 ),
⎪
⎪
⎩
Z = Z0 + h c1 f (g1 ) + (c1 + c2 )f (g2 ) + (c2 + c3 )f (g3 ) + c3 f (g4 ) ,
(1.24)
where g2 = Z1 , g3 = Z2 , g4 = Z3 , and Z = Z3 . Thus, the Butcher tableau
C A
bT
takes the following form:
0 0 0 0 0
2c1 c1 c1 0 0
2(c1 + c2 ) c1 c1 + c2 c2 0
2(c1 + c2 + c3 ) c1 c1 + c2 c2 + c3 c3
c1 c1 + c2 c2 + c3 c3
From the previous chapter, we have the order condition for R–K method as follows:
s
giJ = Z0J + aij f J (gj1 , · · · , gjh ),
j=1
s
Z1J = Z0J + bj hf J (gj1 , · · · , gjh )
j=1
is order of p, iff
s
1
bj φj (ρt) =
j=1
γ(ρt)
for all rooted tree ρt, have r(ρt) ≤ p, where Z0 = (Z01 , · · · , Z0n )T , f J = (f 1 , f 2 , · · · , f n )T .
where ρt is a labelled tree with root j, the sum over the r(ρt) − 1 remaining indices
k, l, · · ·. The summand is a product of r(ρt) − 1 a’s, where all fathers stand two by
two with their sons as indices.
⎧ 4 4
⎪ C
4
⎪
⎪ 1
⎪
⎪ b j = 1, b j ajk = ,
⎪
⎪ j=1 2
⎪
⎪ j=1 k=1
⎪
⎪ 4
4 4
4
⎪
⎪ 1 1
⎪
⎪ bj ajk ajl = , bj ajk akl = ,
⎨ 3 6
j=1 k,l=1 j=1 k,l=1
4 4 4 4 (1.25)
⎪
⎪
⎪
⎪ bj
1
ajk ajl ajm = , bj
1
ajk akl ajm = ,
⎪
⎪
⎪
⎪
4 8
⎪
⎪
j=1 k,l,m=1 j=1 k,l,m=1
⎪
⎪ 4 4 4 4
⎪
⎪ 1 1
⎪
⎩ bj ajk akl akm = , bj ajk akl alm = .
12 24
j=1 k,l,m=1 j=1 k,l,m=1
These Equations (1.26) and (1.27) are just the same as in Subsection 8.1.1 for single
equation.
372 8. Composition Scheme
From the literature[Fen85] , we know that the centered Euler scheme is symplectic,
but trapezoidal scheme (1.1) is right non-symplectic as a result of nonlinear transfor-
mation [Dah75,QZZ95,Fen92] , the scheme (1.1) can transform the Euler center point form,
therefore the trapezoidal form is nonstandard symplectic, just as discussed in Section
4.3 of Chapter 4. It is the same with the trapezoidal form, the centered Euler scheme
may also be used to construct the higher order scheme. Because
⎧
⎪ Z 0 + Z1
⎪
⎪ Z1 = Z0 + d 1 hf ,
⎪
⎪ 2
⎨
Z + Z2
Z2 = Z1 + d2 hf 1 ,
⎪
⎪ 2
⎪
⎪
⎪
⎩ Z3 = Z2 + d3 hf Z2 + Z3 ,
2
it is equally the same in R–K method with in the following Butcher tableau:
d1 d1
0 0
2 2
d2 d2
d1 + d1 0
2 2
d3 d3
d 1 + d2 + d1 d2
2 2
d1 d2 d3
1
Using the same method, we can prove that , when d1 = d3 = 1 , d2 =
2 − 23
1
−2 3
1 , the above scheme is fourth order, and the coefficient is entirely the same as in
2 − 23
trapezoidal method.
⎧ n+ 1 τ
⎪
⎪ p 2 = pn − Vq (q n ),
⎪
⎪ 2
⎪
⎨ q n+ 12 = q n + τ Up (pn+ 12 ),
⎪
⎪
1 τ
pn+1 = pn+ 2 − Vq (q n+ 2 ),
1
⎪
⎪ 2
⎪
⎩ n+1 1
q = q n+ 2 .
This scheme is equal to the following:
⎧ n+ 1 τ
⎪
⎪ p 2 = pn − Vq (q n ),
⎪
⎪ 2
⎪
⎪ τ
⎪ 1
⎨ q n+ 2 = q n +
1
Up (pn+ 2 ),
2
1 τ 1 (2.2)
⎪
⎪ q n+1 = q n+ 2 + Up (pn+ 2 ),
⎪
⎪ 2
⎪
⎪ τ
⎪
⎩ pn+1 = pn+ 2
1
− Vq (q n+1 ).
2
This 2nd order scheme can also be defined as a self-adjoint scheme, see also [Yos90] .
Ruth[Rut83] , using scheme (2.1), constructed a 3rd order scheme via composition
method,
⎧
⎨ p1 = p − c1 τ Vq (q ),
k k
⎪ q1 = q k + d1 τ Up (p1 ),
p2 = p1 − c2 τ Vq (q1 ), q2 = q1 + d2 τ Up (p2 ), (2.3)
⎪
⎩ k+1
p = p2 − c3 τ Vq (q2 ), q k+1
= q2 + d3 τ Up (p k+1
).
7 4 1 2 2
When c1 = , c2 = , c3 = − , d1 = , d2 = − , d3 = 1, this scheme is
24 3 24 3 3
3rd order. We may construct multistage schemes, in order to achieve the higher order
precision. In literature[QZ92,Fen86,Fen91,FR90,Rut83] , we may see the following 4th order
form: ⎧
⎪
⎪ p1 = pk − c1 τ Vq (q k ), q1 = q k + d1 τ Up (p1 ),
⎪
⎪
⎪
⎨ p2 = p1 − c2 τ Vq (q1 ), q2 = q1 + d2 τ Up (p2 ),
(2.4)
⎪
⎪ p3 = p2 − c3 τ Vq (q2 ),
⎪
⎪
q3 = q2 + d3 τ Up (p3 ),
⎪
⎩ k+1
p = p3 − c4 τ Vq (q3 ), q k+1 = q3 + d4 τ Up (pk+1 ),
where ⎧ 1 1
⎪
⎪ c1 = 0, c2 = c4 = − (2 + α), c3 = (1 + 2α),
⎪
⎪ 3 3
⎪
⎨ 1
d1 = d4 = (2 + α),
6 D
⎪
⎪
⎪
⎪ 1 √ 3 1
⎪
⎩ d2 = d3 = (1 − α), α = 2 +
3
,
6 2
or ⎧ 1 1
⎪
⎪ c1 = c4 = (2 + α), c2 = c3 = (1 − α),
⎪
⎪ 6 6
⎪
⎨ 1
d1 = d3 = (1 + 2α),
3 D
⎪
⎪
⎪
⎪ 1 √ 3 1
⎪
⎩ 2
d = − (1 + 2α), d 4 = 0, α = 3
2 + .
3 2
374 8. Composition Scheme
where τ is the step size. S(τ ) is called the integrator, but yn+1 and yn are numerical
solutions of equation in steps n + 1 and n .
Definition 2.1. An integrator S ∗ (τ ) is called the adjoint integrator S(τ ), if
That means yn+1 = S(τ )yn , yn = S ∗ (−τ )yn+1 , or yn+1 = S ∗ (−τ )yn , yn =
S(τ )yn+1 . In fact, (2.6) – (2.7) equations are equivalent to
S(−τ )S ∗ (τ ) = I, (2.8)
S ∗ (τ )S(−τ ) = I. (2.9)
In order to prove this, set τ = −τ , then (2.6) – (2.7) becomes
⎧
⎨ S ∗ (τ )S(−τ ) = I,
⎩ S(−τ )S ∗ (τ ) = I.
τ expresses the length of arbitrary step; therefore, the above equations are formula of
(2.8), (2.9). Further, we would like to point out that the two conditions (2.6) and (2.7)
are the same. Since form S ∗ (−τ )S(τ ) = I, we get
S ∗ (−τ ) = S −1 (τ ),
where S −1 (τ ) is the inverse of the integrator S(τ ). Here, we always assume S(τ ) is
invertible. So, we have
yn , yn+1 is numerical solution for the equation y = f (x) in the n and n + 1 step, and
φ is increment function which the form (2.5) corresponds.
Definition 2.2. Scheme yn+1 = yn + τ φ∗ (x, yn , τ ) is an adjoint scheme (2.10), if it
satisfies:
B = A − τ φ(x + τ, A, −τ ), (2.11)
A = B + τ φ∗ (x, B, τ ). (2.12)
B + τ φ∗ (x, B, τ ) = A,
Since
S ∗ (τ )S(−τ )yn+1 = S ∗ (τ ) yn+1 − τ φ(x + τ, yn+1 , −τ )
= A − τ φ(x + τ, A, −τ ) + τ φ∗ (x, B, τ ),
i.e., S ∗ (τ )S(−τ ) = I.
In[Yos90] , the integrator with the property S(−τ )S(τ ) = I. H. Yoshida called this
operator as reversible. We see that time reversibility and self-adjointness are the same.
The time reversible (i.e., self-adjoint) integrator plays an important role in this chapter
due to its special property.
τ τ τ τ
Theorem 2.5. For every integrator S(τ ), S ∗ S or S S∗ is self-
2 2 2 2
adjoint integrator[QZ92,Str68] .
τ τ
Proof. We must prove that S ∗ =S is self-adjoint, i.e.,
2 2
∗
S ∗ (τ )S(τ ) = S ∗ (τ )S(τ ).
where S1 (τ ) and S2 (τ ) are self-adjoint operators, but they not commutative. We will
construct a higher order form in the below section.
It is linear and satisfies the Leibniz formula, i.e., for two scalar functions ϕ1 and ϕ2 ,
n
∂
n
∂ n
∂fi ∂ n
∂2ϕ
Lg Lf ϕ = g T Df T Dϕ = gj fi ϕ= gj ϕ+ gj fi ,
j=1
∂yj i=1 ∂yi i,j=1
∂yj ∂yi i,j=1
∂yj ∂yi
therefore,
n %
&
∂gi ∂fi ∂
(Lf Lg − Lg Lf )ϕ = fj − gj ϕ,
i,j=1
∂yj ∂yj ∂yi
this means
n %
&
∂gi ∂fi
[Lf , Lg ] = Lc , c = [c1 , c2 , · · · , cn ], ci = fj − gj ,
j=1
∂yj ∂yj
and Lc will still be the first-order differential operator. It is very easy to prove the
following properties of bracket
[Lf , Lf ] = 0, (3.5)
[λ1 Lf1 + λ2 Lf2 , Lg ] = λ1 [Lf1 , Lg ] + λ2 [Lf2 , Lg ], ∀ λ1 , λ2 ∈ R. (3.6)
The commutator also satisfies the Jacobi identity, i.e., if Lf , Lg , Lh are three first-
order differential operators, then
@ A @ A @ A
[Lf , Lg ], Lh + [Lg , Lh ], Lf + [Lh , Lf ], Lg = 0. (3.7)
(3.7) is very easy to prove, the detailed proof process is seen [Arn89] . From the above,
we know that first-order differential operator forms a Lie algebra.
∞ k
* n +k
t d
et[1,1,···,1]D ϕ(y) = ϕ(y)
k ! i=1 dyi
k=0
= ϕ y + t(1, 1, · · · , 1)T . (3.10)
We have seen several properties of Lie series, which are similar to those of [Ste84] . Let,
and T T T T T
etf D
g = etf D g1 , etf D g2 , · · · , etf D gn ,
then, we have the following:
Property 3.3. The Lie series has the compositionality
Since * n +k
∞ k k
∞ k
tLf
t Lf t ∂
e y= y= fi y,
k! k! i=1
∂y i
k=0 k=0
then
*
∞ k ∞ k
+
t k−1 t k−1
tLf
gm (e y) = gm y1 + Lf f1 (y), · · · , yn + L fn (y)
k! k! f
k=1 k=1
∞ k
* n +k
t ∂
= fi (y) gm (y) = etLf gm (y).
k ! i=1 ∂yi
k=0
then
ẏ(t) = f (y(t)).
T
(y)D
Proof. Since yi (t) = etf yi (0), then
d T T
yi (t) = etf (y)D f T (y)Dyi (0) = etf D fi (y).
dt
From Property 3.3, we have
d T
yi (t) = fi (etf D y) = fi y1 (t), y2 (t), · · · , yn (t) = fi (y(t)).
dt
The proof can be obtained.
From Property 3.6, we know that equation ẏ = f (y) can express the solution y(t) =
etLf y(0), Section 8.2 has discussed that the integral S(τ ) can also be represented
in this form. If S(τ ) has the group property about τ , it will be the phase flow of
dy
autonomous ODE = f (y). However, in our problem, there is just one parameter
dτ
family S(τ ) without group property. So, there just exists a formal vector field f τ (y),
which defines only the formal autonomous system
dy
= f τ (y).
dt
Its formal phase flow concerning two parameters τ, t, can be expressed by etLf τ . Take
the diagonal phase flow
8.3 Construction of Higher Order Schemes 381
etLf τ |t=τ = eτ Lf τ .
This is just S(τ ) Lie series expression. See the next chapter to know more about the
formal vector field and the formal phase flow. Since f τ (y) is a formal vector field, it
is a formal power series in τ . Thus, the exponential representation of S(τ ) will the
following form:
and series
τ A + τ 2B + τ 3C + τ 4D + τ 5E + · · ·
may not be convergence, where A, B, C, D, E, · · · are first-order differential opera-
tors. Therefore, we have:
Theorem 3.7. Every integrator S(τ ) has a formal Lie expression [QZ92] .
We use Theorem 3.7 to derive an important conclusion.
Theorem 3.8. Every self-adjoint integrator has an even order of accuracy[QZ92] .
Proof. Let S(τ ) be a self-adjoint integrator. Expand S(τ ) in the exponential form
then n+1
S(τ ) = eτ Lf +O(τ )
.
We must show that n is an even number. This means that we have to prove
w2 = w4 = w6 = w8 = · · · = 0.
Since
S(−τ ) = exp (−τ w1 + τ 2 w2 − τ 3 w3 + · · ·),
and using the BCH formula, we get
Since S(τ ) is self-adjoint, i.e., S(τ )S(−τ ) = I, So (3.14) means w2 = 0, and (3.14)
becomes
S(τ )S(−τ ) = exp 2τ 4 w4 + O(τ 5 ) .
This leads to w4 = 0. Continuing this process, we have
382 8. Composition Scheme
w2 = w4 = w6 = · · · = w2k = · · · = 0.
Therefore, if S(τ ) is of order n, then n must be an even number. Since S(τ ) is at least
of order 1, and if n 2, we have w1 = Lf , because S(τ ) Lie series expression is
unique.
2c2n+1
1 + c2n+1
2 = 0, 2c1 + c2 = 1,
H. Yoshida in [Yos90] obtained the same result for symplectic explicit integrator
used to solve separable systems. The result can be applied to non-Hamiltonian systems
and non-symplectic integrators.
In this chapter, we will extend these results to solve general autonomous system’s
form. Some examples of adjoint scheme and its construction are given below. A con-
crete method to construct an adjoint for any given scheme is also given. This method
can be referred in literature [HNW93] . If the numerical solution is yτ , then any given
scheme may be expressed as:
Lemma 3.10. Every R–K method has an adjoint method, whose coefficients a∗ij , b∗j , c∗j
(i, j = 1, · · · , s) can be written as follows:
⎧ ∗ ∗
⎨ ci = 1 − cs+1−i ,
⎪
a∗ij = bs+1−j − as+1−i,s+1−j , (3.20)
⎪
⎩ ∗
bj = bs+1−j .
Lemma 3.11. If as−i+1,s−j+1 + aij = bs−j+1 = bj , then the corresponding R–K
method (3.18) is self-adjoint.
Concentrating on semi-explicit symplectic R–K method, we have:
Theorem 3.12. The semi-explicit symplectic R–K method for autonomous systems is
self-adjoint if its Butcher tableau is of the form[QZ92] .
Proof. We know that the Butcher tableau of semi-explicit symplectic R–K method
must be of the form
b1
2
b2
b1
2
.. .. ..
. . .
bs−1
b1 b2 ···
2
bs
b1 b2 ··· bs−1
2
b1 b2 ··· bs−1 bs
and S −1 (−τ ) are symplectic integrators, and therefore S ∗ (τ ) is also symplectic. The
two Lemmas given below can be seen in paper[HNW93] . Lemma 3.13 is derived from
Theorem 1.24 of Chapter 7 and Theorem 1.1 of Chapter 8.
Lemma 3.13. If in an implicit s-stage R–K method, all ci (i = 1, · · · , s) are different
and at least of order s, then it is a collocation method iff it is satisfies:
s
cqi
aij cq−1
j = , i = 1, · · · , s, q = 1, · · · , s. (3.21)
j=1
q
where
(wqi , qi )
δi+1 = , i ≥ 0,
(qi , qi )
⎧
⎨ 0, for i = 0,
2
γi+1 = (qi , qi )
⎩ , for i ≥ 1,
(qi−1 , qi−1 )
and - 1
(qi , qj ) = qi (w)qj (w) d w.
−1
1
We obtain q1 = w, q2 = w2 − , assuming q2n (w) is an even function and q2n−1 (w)
3
is an odd function. We proceed by induction on n, for n = 1, this has established.
386 8. Composition Scheme
Suppose q2n is an even function, and q2n−1 is an odd function. Prove that n + 1 is also
true.
Since
- 1
odd d w
(wq2n , q2n ) even|1−1
δ2n+1 = = −1 = = 0,
(q2n , q2n ) (q2n , q2n ) (q2n , q2n )
2
then q2n+1 = wq2n − γ2n+1 q2n−1 (w) = odd function−odd function = odd function.
- 1
odd dw
(wq2n+1 , q2n+1 ) −1
δ2n+2 = = = 0,
(q2n+1 , q2n+1 ) (q2n+1 , q2n+1 )
2
then q2n+2 = wq2n+1 − γ2n+2 q2n (w) = is an even function. We have proved this
conclusion for n + 1. From this, the P2n (w) root may be written in the following
sequence:
−w1 , −w2 , · · · , −wn , wn , · · · , w2 , w1 .
But the root of p2n+1 (w) has the following form:
−w1 , −w2 , · · · , −wn , 0, wn , · · · , w2 , w1 ,
where wi > 0, wi > wi+1 (i = 1, · · · , n), therefore wi + wn+1−i = 0. Even though
we have the direct proof of Theorem 3.15, the computation is tedious. As a result of
Gauss–Legendre method coefficient aij , bj satisfies the following equation:
s
cqi
aij cq−1
j = , i = 1, · · · , s, q = 1, · · · , s, (3.22)
q
j=1
s
1
bj cjq−1 = , q = 1, · · · , s. (3.23)
q
j=1
1
s (−1)n+k
bj = ; k ϕ , i, j, l = 1, · · · , s,
kj
k=1 (cj − cl )
j=l
where
⎧
⎪
⎨ ϕkj = ct1 ct2 · · · cts−k , k < s,
{t1 ,t2 ,···,ts−k }⊂{1,2,···,j−1,j+1,···,s}
⎪
⎩
ϕsj = 1.
8.3 Construction of Higher Order Schemes 387
cki - ci
s (−1)n+k
; k ϕkj = lj (t)d t, j = 1, · · · , s, (3.24)
k=1 (cj − cl ) 0
j=l
s (−1)n+k 1 - 1
; k ϕkj = lj (t)d t, j = 1, · · · , s, (3.25)
k=1 (cj − cl ) 0
j=l
B
(t − ck )
k=j
where lj = B , when ci = 1 − cs+1−i , we have li (t) = ls+1−i(1−t) , then
(cj − ck )
k=j
Composition scheme (2.2) from (2.1) and (3.26) is of order 2. In order to maintain
τ
the transient step size, original τ will shrink and will become , because the present
2
∗ τ τ
scheme is S S . If it is self-adjoint, the transient length of step size is main-
2 2
tained as τ .
Example 3.18. The explicit 4th order symplectic scheme (2.4) can be composed by
S2 (x1 τ ) S2 (x2 τ ) S2 (x1 τ ) and developed as follows:
SV (c1 τ )SU (d1 τ )SV (c2 τ )SU (d2 τ )SV (c3 τ )SU (d3 τ )SV (c4 τ )SU (d4 τ ).
0 12 3 0 12 30 12 3
S2 (x1 τ ) S2 (x2 τ ) S2 (x1 τ )
x1 1
c1 = c4 = = √ = 0.6756035,
2 2− 33
1
d1 = d3 = x1 = √ = 1.35120719,
2(2 − 3 2)
√
−32
d2 = √ = x2 = −1.7024142, d4 = 0,
2− 32
√
x + x2 1− 32
c2 = c3 = 1 = √ = −0.1756036.
2 2(2 − 3 2)
Example 3.19. By literature[FQ91] , we know that one class of symplectic scheme for
dy
equation = J∇H is
dt
% &
1 1
y k+1 = y k + τ J(∇H) (I + JB)y k+1 + (I − JB)y k , B T = B, (3.27)
2 2
5 6 5 6
O −In In O
J= , I= ,
In O O In
this scheme is of order 1, if B = O; if B = O, then the scheme will be of order 2.
−2 3
1
yn+ 23 = yn+ 13 + 1 τ f (yn+ 13 ) + f (yn+ 23 ) , (4.1)
2(2 − 2 ) 3
1
yn+1 = yn+ 23 + 1 τ f (yn+ 23 ) + f (yn+1 ) .
2(2 − 2 ) 3
8.4 Stability Analysis for Composition Scheme 389
We will prove that although the trapezoid method is A-stable, scheme (4.1) is not
A-stable. Fortunately , the unstable region is very small, as Fig. 4.2 (enlaged figure is
Fig.4.1) shows, and scheme (4.1) is still useful for solving stiff systems. Judging from
the size and location of the unstable region of scheme (4.1), we know it is safe for
systems which have eigenvalues not very adjacent to the real axis, while some other
methods which have unstable regions near the imaginary axis, such as Gear’s are safe
for systems which have eigenvalues not very adjacent to the imaginary axis.
0.020
0.012 S
0.004
–0.004
–0.012
–0.020
–0.596 –0.579 –0.563
–0.603 –0.587 –0.571
Fig. 4.1. Closed curve S which contains all zero point of scheme (4.1)
1.0
0.6
0.2
0.0
–2.0 –1.6 –1.2 –0.8 –0.4
–0.2
–0.6
–1.0
Just the same as in scheme (4.1), the Euler midpoint rule can also be used to
construct a scheme:
⎧ % &
⎪ 1 yn + yn+ 1
⎪
⎪ yn+ 13 = yn + 1 τf
3
,
⎪
⎪ 2 − 23 2
⎪
⎪
⎪
⎪ % &
⎨ 1
−2 3 yn+ 1 + yn+ 2
3 3
yn+ 23 = yn+ 13 + 1 τ f , (4.2)
⎪
⎪ 2 − 23 2
⎪
⎪ % &
⎪
⎪ 1 yn+ 2 + yn+1
⎪
⎪ yn+1 = yn+ 23 + 3
⎪ 1 τf .
⎩ 2 − 23 2
390 8. Composition Scheme
Scheme (4.2) is symplectic, but scheme (4.1) is non-symplectic. We now study the
stability of scheme (4.1). Note that scheme (4.1) is not A- stable, whereas the trapezoid
method is. To show this, we apply scheme (4.1) to test equation
which yields
⎧ τ
⎪
⎪ yn+ 13 = yn + c1 λyn + λyn+ 13 ,
⎪
⎪ 2
⎨ τ
yn+ 23 = yn+ 13 + c2 λyn+ 13 + λyn+ 23 , (4.4)
⎪ 2
⎪
⎪ τ
⎪
⎩ n+1
y = y 2
n+ 3 + c1 λyn+ 23 + λyn+1 ,
2
i.e., ⎧
⎪ c1 λτ
⎪
⎪ 1+
⎪
⎪ yn+ 13 = 2 y ,
⎪
⎪ c1 λτ n
⎪
⎪ 1−
⎪
⎪ 2
⎪
⎪
⎪
⎪
⎪
⎨ 1+
c2 λτ
yn+ 3 =
2
2 y 1, (4.5)
⎪
⎪ −
c2 λτ n+ 3
⎪
⎪ 1
⎪
⎪ 2
⎪
⎪
⎪
⎪ c 1 λτ
⎪
⎪ 1+
⎪
⎪ 2 y 2,
⎪
⎪ yn+1 =
c1 λτ n+ 3
⎩ 1−
2
1
1 −2 3 λτ
where c1 = 1 , c2 = 1 . Let = z, z ∈ C, we have
2 − 23 2 − 23 2
(1 + c1 z)(1 + c2 z)(1 + c1 z)
yn+1 = yn . (4.6)
(1 − c1 z)(1 − c2 z)(1 − c1 z)
i.e., 7
(1 + c1 z)(1 + c2 z)(1 + c1 z)
R= z ∈ C < 1, Re z < 0 . (4.7)
(1 − c1 z)(1 − c2 z)(1 − c1 z)
1
Obviously, when z → (< 0), we have
c2
(1 + c1 z)(1 + c2 z)(1 + c1 z)
(1 − c1 z)(1 − c2 z)(1 − c1 z) −→ ∞.
8.4 Stability Analysis for Composition Scheme 391
1
This means schemes (4.1) cannot be stable in the adjacent region . Thus, we obtain
c2
the following theorem:
Theorem 4.2. Scheme (4.1) is not A-stable.
Since scheme (4.1) is not A-stable, we will figure out the stable region of it. To do
this, we will first study the roots of the following equation:
(1 + c1 z)(1 + c2 z)(1 + c1 z)
(1 − c1 z)(1 − c2 z)(1 − c1 z) = 1. (4.8)
Once the roots of (4.8) are known, it is not difficult to get the stable region of (4.1).
Note Equation (4.8) is equivalent to
(1 + c1 z)(1 + c2 z)(1 + c1 z)
= eiθ , 0 ≤ θ < 2π. (4.9)
(1 − c1 z)(1 − c2 z)(1 − c1 z)
Since
392 8. Composition Scheme
2y + w = p1 , (4.15)
x2 + y 2 + 2yw = −p2 , (4.16)
x2 w + y 2 w = −p3 . (4.17)
Now, we will prove that Equations system (4.15)–(4.17) exists as a set of real solution.
In fact, from (4.16) and (4.17), we get:
p2 w + 2yw2 = p3 . (4.18)
w3 − p1 w2 − p2 w + p3 = 0. (4.19)
Since p1 , p2 , p3 are all real, (4.19) is a polynomial with real coefficient and has one
real root and two conjugate complex roots. Using the real root w from (4.19), we can
get a real value of y from (4.15) , and from (4.16) and (4.17), we have x2 = real, then
x is real or pure imaginary. If x is pure imaginary, then from (4.12), z1 , z2 , z3 are all
pure imaginaries, so all the the roots of (4.11) are on the imaginary axis. This means
that if we assume
(1 + c1 z)(1 + c2 z)(1 + c1 z)
V (z) = ,
(1 − c1 z)(1 − c2 z)(1 − c2 z)
then, V (z) > 1 or V (z) < 1 for Re (z) < 0. For the same reason that scheme (4.1)
cannot be A-stable, V (z) > 1 for Re (z) < 0 is possible, but we have V (−0.5) < 1,
1
and V (z) is continuous except at . Thus, x is impossible to be pure imaginary, so
c2
it must be real. Since polynomial (4.11) only has three roots, we will get the same
results of z1 , z2 , z3 if we use other value of w, real or complex.
Theorem 4.5. The three roots of polynomial (4.11) are in the form :
Theorem 4.5 tells us that there is a root of (4.10) on the imaginary, and that the two
other roots are located symmetrically with respect to the imaginary axis. Thus, there
is only one root on the left open semi-plane. Computation shows that these roots form
a closed curve S (when θ changes from 0 to 2π), as in Fig. 4.1.
From (4.15) – (4.17), we get the equation for S:
2
x − 3y 2 + 2p1 y + p2 = 0,
0 ≤ θ < 2π, θ = π. (4.20)
2yx2 + 2y 3 − p1 x2 − p1 y 2 − p3 = 0,
and D
1
x=± − , y = 0, for θ = π, (4.21)
b
b sin θ 1 sin θ
where z = −x + iy, p1 = , p2 = , p3 = .
a(1 + cos θ) a a(1 + cos θ)
Since in S, V (z) > 1, and when z → ∞, V (Z) → 1, we can conclude the
stability of scheme (4.1).
Theorem 4.6. The stable region R of scheme (4.1) is [QZ93] :
Example 4.7. Numerical test for orders of schemes (4.1) and (4.2).
To test the order scheme (4.1) and (4.2), we apply them to the following Hamilto-
nian system: ⎧
⎪ d p = − ∂ H = −w2 q − q 3 ,
⎪
⎨ dt ∂q
(4.22)
⎪
⎪
⎩ d q = ∂ H = p,
dt ∂p
394 8. Composition Scheme
1 1
where the Hamiltonian H = p2 + w2 q 2 + q 4 , and compare the numerical solu-
2 2
tions with trapezoid method and centered Euler scheme.
For convenience, the numerical solution of p and q can be denoted as
1◦ (4.1) by T 4p, T 4q.
2◦ (4.2) by E4p, E4q.
3◦ trapezoid scheme by T 2p, T 2q.
4◦ centered Euler scheme by E2p, E2q. Respectively, we use double precision
in all computations. We can see the following explicit scheme:
⎧ τ
⎪
⎪ p 1 = pn − Vq (qn ),
⎪ n+ 2
⎪ 2
⎪
⎪
⎨ qn+ 1 = qn + τ Up (pn+ 1 ),
2 2 2
(4.23)
⎪ τ
⎪
⎪ qn+1 = qn+ 12 + 2 Up (pn+ 12 ),
⎪
⎪
⎪
⎩ p τ
= p 1 − V (q
n+1 n+ 2 ), q n+1
2
x1 1
c1 = c4 = = √ = 0.6756035,
2 2− 33
1
d1 = d3 = x1 = √ = 1.35120719,
2(2 − 3 2)
√
−32
d2 = √ = x2 = −1.7024142, d4 = 0,
2− 32
√
x + x2 1− 32
c2 = c3 = 1 = √ = −0.1756036.
2 2(2 − 3 2)
A separable system with H = V (q) + U (p) is self-adjoint, so it can be used
to construct fourth-order scheme to get (4.1) and (4.2). From Sections 8.2 and 8.3,
1
the simplified fourth-order scheme can be written taking c1 = 1 = c3 , c2 =
2 − 23
1
−2 3
1 , x1 = x3 = 1.35120719, x2 = −1.7024142. For details see Example 3.18.
2 − 23
⎧ τ
⎪
⎪ pn+ 14 = pn − x1 Vq (qn ),
⎪
⎪ 2
⎪
⎪
⎪
⎪ q 1 = qn + x1 τ Up (p
n+ 14 ),
⎪
⎪
n+ 3
⎪
⎪ x + x
⎪
⎪ pn+ 12 = pn+ 14 − 1 2
τ Vq (qn+ 13 ),
⎪
⎪ 2
⎨
qn+ 23 = qn+ 13 + x2 τ Up (pn+ 12 ), (4.24)
⎪
⎪
⎪
⎪ x + x3
pn+ 34 = pn+ 12 − 2
⎪
⎪ τ Vq (qn+ 23 ),
⎪
⎪ 2
⎪
⎪
⎪
⎪ qn+1 = qn+ 23 + x3 τ Up (pn+ 34 ),
⎪
⎪
⎪
⎪ τ
⎩ pn+1 = pn+ 3 − x3 Vq (qn+1 ),
4 2
where pn+ 14 , pn+ 12 , pn+ 34 and qn+ 13 , qn+ 23 denote the numerical solution of different
stages at every step. Scheme (4.24) has been proved by H. Yoshida to be a fourth-
order scheme in [Yos90] . We can apply scheme (4.24) to Equation (4.22) and compare
8.4 Stability Analysis for Composition Scheme 395
the results with that of schemes we mentioned above. We denote EXp and EXq as
the exact solution of p and q for system (4.22), and present our results when taking
w = 2, τ = 0.1, p0 = 0.5, q0 = 0.5 in Table 4.1. From Table 4.1, we can see that
T 4p, T 4q and E4p, E4q are more approximate to EXp, EXq than T 2p, T 2q and
E2p, E2q. Thus, we conclude that scheme (4.1) and (4.2) have a higher order than
trapezoid method and centered Euler scheme. Table 4.1 also shows that although the
trapezoid scheme (4.1) is non-symplectic, it can be used to solve a Hamiltonian system
to get satisfactory results than the centered Euler scheme, by nonlinear transformation;
the latter can be obtained from the former, see Section 8.1.
Example 4.8. Numerical test for stability of schemes (4.1) and (4.2). To consider
the unstable case, we take λ = −11.8, τ = 0.1, and initial value y0 = 1.0 in the
test equation, so λτ falls into the unstable region. While the exact solution decreases
quickly, the numerical solution obtained by scheme (4.1) grows to infinity as shown
in Table 4.2.
Example 4.9. For the stable case, we consider a linear stiff system
ẏ1 = −501y1 + 500y2 ,
(4.25)
ẏ2 = 500y1 − 501y2 ,
⎧
⎨ y1 (t) = f1 (y1 , y2 ) = 0.5 y1 (0) − y2 (0) e−1001t + 0.5 y1 (0) + y2 (0) e−t ,
⎩ y2 (t) = f2 (y1 , y2 ) = −0.5 y1 (0) − y2 (0) e−1001t + 0.5 y1 (0) + y2 (0) e−t ,
(4.26)
where y1 (0), y2 (0) denote the initial value. Since system (4.25) is linear, schemes
(4.1) and (4.2) are equivalent in this case. We present a numerical solution using
scheme (4.1) here. In Table 4.3, we denote the numerical solution of y1 and y2 using
(4.1) by T 4Y 1, T 4Y 2, and the exact solution of y1 and y2 by EXY 1 and EXY 2. We
also assume τ = 0.1, y1 (0) = 1.5, y2 (0) = 0.5, in the Table 4.3, while τ = 0.0005
in the first 50 steps, and τ = 0.1 in the remaining steps.
neglected. In this section, we use scheme (4.1) to solve two kinds of PDEs in order to
show that the technique introduced in previous section can be used to overcome the
deficiency in the time direction, since theoretically, we can construct arbitrary even
order schemes in the time direction[ZQ93b] .
Let us first consider the following one-dimensional first-order wave equation
ut + ux = 0,
(5.1)
u(x, 0) = f (x), 0 ≤ x ≤ 2π,
Since collocation, Galerkin, and tau methods are identical in the absence of essen-
tial boundary conditions, we will analyze the Fourier collocation or pseudospectral
method. Let us introduce the collocation points xn = 2πn/2N (n = 0, · · · , 2N − 1),
and let u = (u0 , · · · , u2N −1 ), where un = u(xn , t). The collocation equation that
approximates (5.1) is as follows:
∂u
= C −1 DCu, (5.2)
∂t
1 @ A
ckl = √ exp (k − N )xl , (5.3)
2N
dkl = −k ∗ δkl , (5.4)
Table 5.1. Comparison between numerical and exact solution when τ =0.1
Step N n EX ORD.4 ORD.2 ERR.4 ERR.2
0 −0.099833416647 −0.099832763924 −0.099750623437 0.000000652723 0.000082793209
N =1 5 0.099833416647 0.099832763924 0.099750623438 −0.000000652723 −0.000082793209
9 −0.665615704994 −0.66561545443 −0.665553604585 0.000000489551 0.000062100409
0 −0.841470984808 −0.841467440655 −0.841021115481 0.000003544153 0.000449869327
N = 10 5 0.841470984808 0.841467440655 0.841021115481 −0.000003544153 −0.000449869327
9 −0.998346054152 −0.998346431587 −0.998393545150 −0.000000377435 −0.000047490998
0 0.544021110889 0.543966068061 0.537020563223 −0.000055042829 −0.007000547666
N = 100 5 −0.544021110889 −0.543966068061 −0.537020563223 0.000055042829 0.007000547666
9 0.933316194418 0.933292641025 0.930296266090 −0.000023553213 −0.003019928328
Table 5.2. Comparison between numerical and exact solution when τ = 0.01
Step N n EX ORD.4 ORD.2 ERR.4 ERR.2
0 0.009999833340 −0.099998333280 −0.009999750000 0.000000000007 0.000000083334
N =1 5 0.009999833340 0.009999833280 0.009999750000 −0.000000000007 −0.000000083334
9 −0.595845898383 −0.595845898378 −0.595845831454 0.000000000005 0.000000066929
0 −0.099833416647 −0.099833416582 −0.099832587427 0.000000000065 0.000000829220
N = 10 5 0.099833416647 0.099833416582 0.099832587427 0.000000000042 −0.000000829220
9 −0.665615704994 −0.665615704952 −0.665615083044 0.000000000003 0.000000621950
0 −0.841470984808 −0.841470984547 −0.841466481987 0.000000000261 −0.000004502821
N = 100 5 0.841470984808 0.841470984547 0.841466481987 −0.000000000267 −0.000004502871
9 −0.998346054152 −0.998346054304 −0.998346533230 −0.000000000152 −0.000000479078
Similarly, in 2nd order PDE, the result of the 4th order scheme is more precise
when compared to the result of the 2nd order scheme in 2 - 4 precision.
Let us take the second order heat conductivity equation
⎧
⎪ ∂u(x, t) ∂ 2 u(x, t)
⎪
⎨ = , 0 < x < π, t ≥ 0,
∂t ∂x2
u(0, t) = u(π, t) = 0, t > 0, (5.5)
⎪
⎪
⎩
u(x, 0) = f (x), 0 ≤ x ≤ π.
N
uN (x, t) = an (t) sin nx, (5.6)
n=1
and ⎧
⎪ d an 2
⎨ d t = −n an ,
⎪
- π (5.7)
⎪ 2
⎪
⎩ an (0) = π f (x) sin nx d x.
0
8.5 Application of Composition Schemes to PDE 399
Table 5.3. Comparison between numerical and exact solution when τ =0.1
Step N n EX ORD.4 ORD.2 ERR.4 ERR.2
1 0.531850090044 0.5318500444815 0.531805704455 0.0000003547710 −0.000044385589
N =1 2 0.860551522611 0.8605520966420 0.860479705219 0.0000005740310 −0.000071817391
3 0.860551522611 0.8605520966420 0.860479705219 0.0000005740310 −0.000071817391
4 0.531850090044 0.5318504448150 0.531805704455 0.0000003547710 −0.000443855890
1 0.216234110142 0.2162355525360 0.216053719560 0.0000001442394 −0.000180390582
N = 10 2 0.349814139737 0.3498764735800 0.349582261644 0.0000023338430 −0.000291878093
3 0.349814139737 0.3498764735800 0.349582269644 0.0000123338430 −0.000291878093
4 0.216234110142 0.2162355522536 0.216053719560 0.0000014423940 −0.000180390582
1 0.003960465877 0.0039605979700 0.003943973573 0.0000001320940 −0.000164923040
N = 50 2 0.006408168400 0.0064083821320 0.006381483292 0.0000002137320 −0.000026685108
3 0.006408168400 0.0064083821320 0.006381483292 0.0000002137320 −0.000026685108
4 0.003960465877 0.0039605979700 0.003943973573 0.0000001320940 −0.000164923040
πj
The initial value of an can be represented in another form. Let xj = (j =
N +1
1, · · · , N ) be collocation points, from collocation equation
N
πjn
an sin = u(xj ), j = 1, · · · , N, (5.8)
n=1
N +1
2
N
πjn
an = u(xj ) sin , n = 1, · · · , N.
N +1 N +1
j=1
2
N
πjn
an (0) = f (xj ) sin , n = 1, · · · , N. (5.9)
N +1 N +1
j=1
The exact solution for Equation (5.5) with boundary condition f (x) = sin x is
e−t sin x. In Table 5.3 and Table 5.4, all symbols carry the same significance. We
take N = 4 for computation.
ut = LN u, (5.10)
Table 5.4. Comparison between numerical and exact solution when τ =0.01
Step N n EX ORD.4 ORD.2 ERR.4 ERR.2
1 0.581936691312 0.581936691316 0.581936642817 −0.000000000004 −0.000000048495
N =1 2 0.941593345844 0.941593345850 0.941593267377 0.000000000006 −0.000000078467
3 0.941593345844 0.941593345850 0.941593267377 0.000000000006 −0.000000078467
4 0.581936693120 0.581936691316 0.581936642817 −0.000000000004 −0.000000048495
1 0.216234110142 0.216234110285 0.216232308172 0.0000000001430 −0.000001801970
N = 100 2 0.349874139137 0.349874139969 0.349811224088 0.0000000002310 −0.000002915049
3 0.349874139137 0.349874139969 0.349811224088 0.0000000000231 −0.000002915049
4 0.276234110142 0.216234110285 0.216232308172 0.0000000001430 −0.000001801970
1
un+1 − un = Δt LN un+1 + LN un . (5.12)
2
which is self-adjoint and of order 2. We can construct a fourth-order scheme by com-
position
1
un+1/3 = un + Δt LN un + LN un+1/3 ,
2(2 − 2
1/3 )
2 1/3
un+2/3 = un+1/3 − Δt LN un+1/3 + LN un+2/3 , (5.13)
2(2 − 21/3 )
1
un+1 = un+2/3 + Δt LN un+2/3 + LN un+1 .
2(2 − 2 )
1/3
Finally, we can point out that scheme (5.13) is unstable for some special step size of
t. Since the diameter of the unstable region is very small, we can always avoid taking
those step-size Δt which make λΔt (λ denotes the eigenvalue of the system to be
solved) fall into the unstable region. Fig. 5.1 shows the solution of the heat equation
when we use scheme (5.13) to solve the (5.11) We take Δt = 0.0097 and N = 24. We
can conclude that while the Crank–Nicolson remains stable, the scheme (5.13) does
not, and solution tends to overflow. For a Detailed numerical test about this problem,
see[ZQ93b] .
Fig. 5.1. Stability comparison between schemes of Crank–Nicolson (L), (5.13) (M) and exact
solution (R) of the heat equation
8.6 H-Stability of Hamiltonian System 401
dz
= Lz, L = JA ∈ sp(2n), H = (z, Az), AT = A, (6.1)
dt
a linear symplectic algorithm
z k+1 = gH
t
(z k ) = G(s, A)z k , k0 (6.2)
Theorem 6.1. Linear symplectic method (6.1) is stable iff the eigenvalues of G(s) are
unimodular and their elementary divisors are linear [Wan94] .
dz
= αJz, α ∈ R, (6.5)
dt
with
α T α
H(z) = H(p, q) = z z = (p2 + q 2 ), A = αI.
2 2
402 8. Composition Scheme
Remark 6.3. It is reasonable to choose (6.5) as the model equation because any linear
Hamiltonian system may turn into the standard form
1
n
H(p, q) = αi (p2i + qi2 ).
2 i=1
Example 6.5. Applying the centered Euler scheme to the test system (6.5), it becomes
1
z = z + μJ(
z + z), μ = αs,
2
−1
1 1
z = I + μJ I − μJ z,
2 2
where
8.6 H-Stability of Hamiltonian System 403
⎡ 1
⎤
1 − μ2 −μ
1 ⎢ 4 ⎥
G(μ) = 1 ⎣ ⎦, (6.9)
1 + μ2 1
4 μ 1 − μ2
4
therefore
⎛ 1 ⎞
% &2 1 − μ2
a1 + a4
=⎝ 4 ⎠ < 1,
1
∀μ = 0.
2 1 + μ2
4
By Lemma 6.4, we know that the centered Euler scheme to all μ = 0 is stable, cer-
tainly it is also stable for μ = 0, therefore, the centered Euler scheme is H-stable.
For the stability region of certain explicit scheme, see the literature [Wan94,QZ90] .
In Section 8.2, we have constructed schemes of difference from 1st order to 4th
order. We will now discusses its stability by applying these schemes to the model
Equation (6.5), we get
z k+1 = Gi (μ)z k , μ = αs, i = 1, 2, 3, 4
Gi is the step transition equation.
* +
1 −μ
G1 (μ) = ,
μ 1 − μ2
⎛ 1 ⎞
1 − μ2 −μ
⎜ 2 ⎟
G2 (μ) = ⎝ ⎠,
1 1
μ 1 − μ2 1 − μ2
4 2
⎛ ⎞
1 1 1 7
1 − μ2 + μ4 −μ 1 − μ2 + μ4
⎜ 2 72 6 1728 ⎟
G3 (μ) = ⎝ ⎠,
1 2 1 1 5 7
μ 1 − μ + μ4 1 − μ2 + μ4 − μ6
6 72 2 72 1728
* +
a1 a2
G4 (μ) = ,
a3 a4
1 1 4 1
a1 = 1 − μ2 + μ + (1 + β)2 μ6 ,
2 24 144
1 1
a2 = −μ 1 − μ2 − (2 + β)(1 + 2β)2 μ4 ,
6 216
1 1 1
a3 = μ 1 − μ2 − (2 + β)(1 − β)μ4 + (2 + β)(1 + β 2 )μ6 ,
6 216 864
1 1 4 1
a4 = 1 − μ2 + μ + (1 + β)2 μ6 .
2 24 144
404 8. Composition Scheme
Theorem 6.6. From the explicit scheme above, the H-stability intervals are (−2, 2),
(−2, 2), (−2.507, 2.507) and (−1.573, 1.573).
Proof. Proof of this theorem can be found in paper of Daoliu Wang[Wan94] and paper
of Mengzhao Qin and Meiqing Zhang[QZ90] .
Bibliography
[MSSS97] A. Murua, J. M. Sanz-Serna, and R. D. Skeel: Order conditions for numerical inte-
grators obtained by composing simpler methods. Technical Report 1997/7, Departemento
de Matemática Aplicada y Computatión, Universidad de Valladolid, Spain, (1997).
[Mur97] A. Murua: On order conditions for partitioned symplectic methods. SIAM J. Numer.
Anal., 34:2204–2211, (1997).
[Mur99] A. Murua: Formal series and numerical integrators, part I: Systems of ODEs and
symplectic integrators. Appl. Numer. Math., 29:221–251, (1999).
[Mur06] A. Murua: The Hopf algebra of rooted trees, free Lie argebra,and Lie series. Founda-
tions of Computational Mathematics, 6(4):387–426, (2006).
[Ner87] F. Neri: Lie algebras and canonical integration. University of Maryland Tech. report,
(1987).
[QWZ91] M. Z. Qin, D. L. Wang, and M. Q. Zhang: Explicit symplectic difference schemes
for separable Hamiltonian systems. J. Comput. Math., 9(3):211–221, (1991).
[QZ90] M. Z. Qin and M. Q. Zhang: Multi-stage symplectic schemes of two kinds of Hamil-
tonian systems for wave equations. Computers Math. Applic., 19:51–62, (1990).
[QZ90a] M. Z. Qin and M. Q. Zhang: Explicit Runge–Kutta–like schemes to solve certain
quantum operator equations of motion. J. Stat. Phys., 60(5/6):839–843, (1990).
[QZ92] M. Z. Qin and W. J. Zhu: Construction of higher order symplectic schemes by com-
position. Computing, 47:309–321, (1992).
[QZ93] M. Z. Qin and W. J. Zhu: A note on stability of three stage difference schemes for
ODEs. Computers Math. Applic., 25:35–44, (1993).
[QZZ95] M. Z. Qin, W. J. Zhu, and M. Q. Zhang: Construction of symplectic of a three stage
difference scheme for ODEs. J. Comput. Math., 13:206–210, (1995).
[Rut83] R. Ruth: A canonical integration technique. IEEE Trans. Nucl. Sci., 30:26–69, (1983).
[Ste84] S. Steinberg: Lie series and nonlinear ordinary equations. J. of Math. Anal. and Appl.,
101:39–63, (1984).
[Str68] G. Strang: On the construction and comparison of difference schemes. SIAM J. Numer.
Anal., 5:506–517, (1968).
[Suz77] M. Suzuki: On the convergence of exponential operators the zassenhuas formula,
BCH formula and systematic approximants. Communications in Mathematical Physics,
57:193–200, (1977).
[Suz90] M. Suzuki: Fractal decomposition of exponential operators with applications to many-
body theories and Monte Carlo simulations. Physics Letters A, 146:319–323, (1990).
[Suz92] M. Suzuki: General theory of higher-order decomposition of exponential operators
and symplectic integrators. Physics Letters A, 165:387–395, (1992).
[Wan94] D. L. Wang: Some acpects of Hamiltonian systems and symplectic defference meth-
ods. Physica D, 73:1–16, (1994).
[Wru96] O. Wrubel: Qin-Kompositionen mit Lie-Reihen. Diplomarbeit Uni Karlsruhe (TH),
(1996).
[Yos90] H. Yoshida: Construction of higher order symplectic integrators. Physics Letters A,
150:262–268, (1990).
[ZQ93] W. Zhu and M. Qin: Applicatin of higer order self-adjoint schemes of PDEs. Comput-
ers Math. Applic., 26(3):15–26, 1993.
Chapter 9.
Formal Power Series and B-Series
We study vector fields, their associated dynamical systems and phase flows together
with their algorithmic approximations in RN from the formal power series approach
[Fen93a,Fen92]
.
9.1 Notation
Our considerations will be local in both space and time, all related objects are C∞
smooth. We use coordinate description and matrix notation, the coordinate vec-
tors in RN and vector functions a : RN → RN are denoted by column matri-
ces. The identity vector function 1N is given by 1N (x) = x. For vector function
a = (a1 , · · · , aN )T : RN → RN ,
∂ ai
a∗ : = = Jacobian matrix a,
∂ xj
∂
a∗ : = ai = linear differential operator of first order associated to a,
∂xi
a∗ b = a∗ (b1 , · · · , bN )T = (a∗ b1 , · · · , a∗ bN )T = b∗ a, a∗ 1N = a.
The Lie algebra VN is associated with the (∞-dimensional) local Lie group DN of
near-identity diffeomorphisms—or simply near-1 maps—of RN .
Consider the dynamical system in RN
dx
= a(x), (1.1)
dt
defined by a vector field a. It possesses a phase flow eta = et , which is a one-parameter
(in t) group of near-1 maps of RN ,
e0 = 1N , et+s = et ◦ es ,
and generates the solution by x(0) → eta x(0) = x(t). The phase flow is expressible
as a convergent power series in t:
∞
eta = 1N + tk ek ,
k=1
1 ∗ 1 1
e0 = 1N , ek = a ek−1 = (a∗ )k 1N = ak .
k k! k!
We define
∞
1
Exp ta∗ := I + (ta)∗k , I is the identity operator.
k!
k=1
This is an operator power series operating on scalar and vector functions, and defined
by
∞
∞ k
∗ 1 ∗ k t k
exp ta := (Exp ta )1N = 1N + (ta ) 1N = 1N + a , (1.2)
k! k!
k=1 k=1
then
eta = (Exp ta∗ )1N = exp ta, (1.3)
for scalar function
φ ◦ eta = φ ◦ exp ta = (Exp ta∗ )φ,
for vector function
Each numerical algorithm solving the system (1.1) possesses the step transition map
fas which is one-parameter (in step-size s) family (in general not a one-parameter
group in s) of near-1 maps on RN , expressible as a convergent power series in s
9.2 Near-0 and Near-1 Formal Power Series 409
∞
fas = 1N + sk fk , (1.4)
k=1
the coefficients can be determined recursively form the defining difference equation.
The transition generates the numerical solution x(0) → (fas )N x(0) ≈ x(N s) by
iterations with step-size s chosen fixed in general.
The main
problem is to construct and analyze the algorithmic approximations
fas ≈ eta t=s = esa in a proper way. For this purpose, we propose a unified frame-
work based on the apparatus of formal power series, Lie algebra of vector fields, and
the corresponding Lie group of diffeomorphisms [Lie88,Olv93] .
The associated near-0 formal differential operators and their products are
* ∞
+ ∞
k
s
(a )∗ : = s ak := sk ak∗ ,
k=1 ∗ k=1
* ∞
+∗ ∞
as∗ := sk ak := sk a∗k ,
k=1 k=1
∞
as∗ bs∗ : = sk a∗i b∗j , (as∗ )2 : = as∗ as∗ , etc.
k=2 i+j=k
∞
(a ◦ g s )(x) = a(g s (x)) = a(x) + sk (a ◦ g)k (x),
k=1
k 1
(a ◦ g)k = (Dm a)(gk1 , · · · , gkm ),
m=1 k1 +···+km =k
m!
where
Dm a = (Dm a1 , · · · , Dm aN )T ,
N
∂ m ai
Dm ai (v1 , · · · , vm ) = v1j · · · vmjm ,
j1 ,···,jm =1
∂xj1 · · · ∂xjm 1
is the usual m-th differential multi-linear form for m tangent vectors vi = (vi1 , · · ·,
viN )T (i = 1, · · · , m) at point x ∈ RN , which is invariant under permutation of
vectors. Using the identities,
(D1 a)(b) = b∗ a,
(D2 a)(b, c) = (c∗ b∗ − (c∗ b)∗ )a,
(D3 a)(b, b, b) = (b∗3 + 2b3∗ − 3b∗ b2∗ )a.
We get in particular
(a ◦ g)1 = g1∗ a,
1
(a ◦ g)2 = g2∗ a + (g1∗2 − g12∗ )a,
2
1 ∗3
(a ◦ g)3 = g3∗ a + ((g2∗ g1∗ − (g2∗ g1 )∗ )a + (g + 2g13∗ − 3g1∗ g12∗ )a.
3! 1
∞
(f s ◦ g s )(x) = f s (g s (x)) = 1N (g s (x)) + sk fk (g s (x))
k=1
∞
=: 1N (x) + sk (f ◦ g)k (x),
k=1
(f ◦ g)1 = f1 + g1 ,
(f ◦ g)k = fk + gk + δ(f1 , · · · , fk−1 ; g1 , · · · , gk−1 ), k ≥ 2,
k−1
i 1
δ(f1 , · · · , fk−1 ; g1 , · · · , gk−1 ) = (Dm fk−i )(gi1 , · · · , gim ).
i=1 m=1 i1 +···+im =i
m!
9.2 Near-0 and Near-1 Formal Power Series 411
In particular we get,
(f ◦ g)2 = f2 + g2 + g1∗ f1 ,
1
(f ◦ g)3 = f3 + g3 + g1∗ f2 + g2∗ f1 + (g1∗2 − g12∗ )f1 ,
2
1
(f ◦ g)4 = f4 + g4 + g1∗ f3 + g2∗ f2 + g3∗ f1 + (g1∗2 − g12∗ )f2
2
1 ∗3
+(g2∗ g1∗ − (g2∗ g1 )∗ )f1 + (g + 2g13∗ − 3g1∗ g12∗ )f1 .
3! 1
form a (∞-dim) formal Lie group FDN . In group FDN , inverse elements, square
roots, rational powers, etc., always exist, and their coefficients can always be deter-
mined recursively by the defining composition relations. For example, the inverse
∞
(f s )−1 := 1 + sk hk = hs is defined by (f s ◦ hs ) = 1N , hence
k=1
In particular,
There is an obvious one-one correspondence between the Lie algebra FVN and the
Lie group FDN , established simply by +1N and −1N . However, the more significant
one-one correspondence between them is given by exp and its inverse log.
Note that
* ∞
+ * ∞
+ ∞
s∗ m
(a ) = sk1 a∗k1 ··· skm a∗km = sk1 +···+km a∗k1 · · · a∗km ,
k1 =1 km =1 k1 ,···,km =1
412 9. Formal Power Series and B-Series
so we get easily
k
1
fk = a∗k1 · · · a∗km 1N , k ≥ 1, f1 = a1 ,
m!
m=1 k1 +···+km =k
k
1 1
fk = ak + a∗k1 · · · a∗km 1N , k ≥ 2, f2 = a2 + a21 . (2.2)
m! 2
m=2 k1 +···+km =k
In particular,
1 1 1 3
a1 = f1 , a2 = f2 − a21 , a3 = f3 − (a∗1 a2 + a∗2 a1 ) − a ,
2 2 3! 1
1 1 ∗ ∗ 1
a4 = f4 − (a∗1 a3 + a22 + a∗3 a1 ) − (a a a2 + a∗1 a∗2 a1 + a∗2 a∗1 a1 ) − a41 ,
2 3! 1 1 4!
k−1
1 1 k
ak = fk − a∗k1 · · · a∗km 1N − a , k ≥ 3.
m! k! 1
m=2 k1 +···+km =k
where
hs1 = f s − 1N , hsm = hsm−1 ◦ f s − hsm−1 .
It is easy to compute
∞
∞
hs1 = sk fk = sk1 (1N ◦ f )k1 ,
k=1 k1 =1
∞
hs2 = sk1 +k2 ((1N ◦ f )k1 ◦ f )k2 ,
k1 ,k2 =1
∞
hs3 = sk1 +k2 +k3 (((1N ◦ f )k1 ◦ f )k2 ◦ f )k3 ,
k1 ,k2 ,k3 =1
···.
9.2 Near-0 and Near-1 Formal Power Series 413
∞
Substituting in (2.3) and equating with sk ak , we get
k=1
k
(−1)m−1
ak = (· · · ((1N ◦ f )k1 ◦ f )k2 · · · ◦ f )km . (2.4)
m
m=1 k1 +···+km =k
It is easy to verify log exp as = as for this log, so this is precisely the inverse of exp,
thus agreeing with the previous one.
We use the above construction (2.4) to establish the formal Baker–Campbell–
Hausdorff formula[Bak05,Hau06] . For arbitrary near-1 formal maps f s , g s ,
∞
log (f s ◦ g s ) = log f s + log g s + dk (log f s , log g s ),
k=1
1 (−1)m−1
k
[(as )p1 (bs )q1 · · · (as )pm (bs )qm ]
dk (as , bs ) = ,
k m p1 !q1 ! · · · pm !qm !
m=1 p1 +q1 +···+pm +qm =k
pi +qi ≥1,pi ≥0,qi ≥0
where
(x)p = xx · · · x (p times), [x1 x2 x3 · · · xn ] = [[· · · [[x1 , x2 ], x3 ], · · ·], xn ].
In particular,
1 s s 1 s s s 1 s s s s
d1 = [a , b ], d2 = [a b b ] + [bs as as ] , d3 = − [a b b a ].
2 12 24
∞
Let log (f s ◦ g s ) = cs = sk ck , then
k=1
1
c1 = a1 + b1 , c2 = a2 + b2 + [a1 b1 ],
2
1 1
c3 = a3 + b3 + ([a1 b2 ] + [a2 b1 ]) + ([a1 b1 b1 ] + [b1 a1 a1 ]),
2 12
1
c4 = a4 + b4 + ([a1 b3 ] + [a2 b2 ] + [a3 b1 ])
12
1
+ ([a1 b1 b2 ] + [a1 b2 b1 ] + [a2 b1 b1 ] + [b1 a1 a2 ] + [b1 a2 a1 ] + [b2 a1 a1 ])
12
1
− [a1 b1 b1 a1 ], etc.
24
Note that the classical BCH formula is restricted to the composition of two one-
parameter groups, where log f s = sa1 and log g s = sb1 .
The log transform reduces matters at the Lie group level to those at the easier level
of Lie algebra. All properties of near-1 formal maps have their logarithmic interpreta-
tions.
414 9. Formal Power Series and B-Series
[Fen93a,Fen92,Fen93b]
C
∞
Proposition 2.1. We list some of them, let log f s = as = sk ak :
k=1
1◦ f s is a phase flow, i.e., f s+t = f s ◦ f t ⇔ log f s = sa1 .
2◦ f s is revertible, i.e., f s ◦ f −s = 1N ⇔ log f s is odd in s.
3◦ f s raised to real μ-th power (f s )μ ⇔ log (f s )μ = μ log f s . In particular,
√ 1
log (f s )−1 = − log f s , log f s = log f s .
2
4◦ f s scaled to f αs ⇔ log (f αs ) = (log f )αs . In particular,
log (f −s ) = (log f )−s .
5◦ f s − g s = O(sp+1 ) ⇔ log f s − log g s = O(sp+1 ).
6◦ f s ◦g s = g s ◦f s ⇔ [log f s , log g s ] = 0 ⇔ log (f s ◦g s ) = log f s +log g s .
C
∞
7◦ (f s ◦g s ) = hs ⇔ log hs = log (f s ◦g s ) = log f s +log g s + dk (log f s , log g s ).
k=1
8◦ f s symplectic ⇔ all ak are Hamiltonian fields (see Chapter 5).
9◦ f s contact ⇔ all ak are contact fields (see Chapter 11).
10◦ f s volume-preserving ⇔ all ak are source-free fields (see Chapter 10).
The log transform has important bearing on dynamical systems with Lie algebra
structure. The structure-preserving property of maps f s at the Lie group (G ⊂ Dm )
level can be characterized through their logarithms at the associated Lie algebra (L ⊂
Vm ) level.
ak
ek = .
k!
fk = ek , 1 ≤ k ≤ p; fp+1 = ep+1 ⇐⇒
a = a1 = e1 ; ak = 0, 1 < k ≤ p; ap+1 = fp+1 − ep+1 = 0. (3.1)
So, the orders of approximation for fas ≈ esa and for log fas − sa are the same.
Moreover, note that we have a formal field
9.3 Algorithmic Approximations to Phase Flows 415
∞
s−1 log f s = s−1 as = a + sk+1 ak+1 = a + O(sp ),
k=1
which is the original field a up to a near-0 perturbation and defines a formal dynamical
system
∞
dx
= (s−1 log f s )(x) = a(x) + sk+1 ak+1 (x)
dt
k=1
having a formal phase flow (in two parameters t and s with group property in t)
ets−1 as = exp ts−1 as whose diagonal formal flow ets−1 as |t=s is exactly f s . This
means that any compatible algorithm fas of order p gives perturbed solution of a right
equation with field a; however, it gives the right solution of a perturbed equation with
field s−1 log fas = a + O(sp ). There could be many methods with the same formal or-
der of accuracy but with quite different qualitative behavior. The problem is to choose
among them those leading to allowable perturbations in the equation. For systems
with geometric structure, the 8◦ , 9◦ , 10◦ of Proposition 2.1 provide guidelines for a
proper choice. The structure-preservation requirement for the algorithms precludes all
unallowable perturbations alien to the pertinent type of dynamics. Take, for example,
Hamiltonian systems. A transition map fas for Hamiltonian field a is symplectic if and
only if all fields ak are Hamiltonian, i.e., the induced perturbations in the equation
are Hamiltonian. So symplectic algorithms are clean, inherently free from all kinds
of perturbations alien to Hamiltonian dynamics (such as artificial dissipation inherent
in the vast majority of conventional methods), this accounts for their superior perfor-
mance. The situations are the same for contact and volume-preserving algorithms .
The Proposition 2.1 profound impact on later developed called “Backward error se-
ries” work, “Modified equation” and “Modified integrator”[Hai94,CHV05,CHV07] .
x1 − x0 = sa(x0 ),
f s − 1N = sa,
fEs = 1N + sa,
s2 2
log fEs = sa − a + O(s3 ),
2
non-revertible, order = 1.
(2) Implicit Euler method (I):
416 9. Formal Power Series and B-Series
x1 − x0 = sa(x1 ),
f s − 1N = sa ◦ f s ,
fIs = (1N − sa)−1 = (fE−s )−1 = 1 + sa + s2 a2 + O(s3 ),
s2 2
log fIs = sa + a + O(s3 ),
2
non-revertible, order = 1.
(3) Trapezoidal method (T ):
s
x1 − x0 = (a(x1 ) + a(x0 )),
2
s
f s − 1N = (a ◦ f s + a),
2
s −1 s s s
fTs = 1N − a ◦ 1N + a = fI2 ◦ fE2
2 2
s s
s2 2 s3
= (fE2 )−1 ◦ fCs ◦ fE2 = 1N + sa + a + a3 + O(s4 ),
2 4
s3 3
log fTs = sa + a + O(s5 ),
12
revertible, order = 2, symplectic for linear Hamiltonian but non-symplectic for non-
linear Hamiltonian systems, where fCs denoting following centered Euler scheme.
(4) Centered Euler method (C):
1
x1 − x0 = sa (x1 + x0 ) ,
2
1 s
f s − 1N = sa ◦ (f + 1N ) ,
2
are conjugate to each other. C is far less known than T , it becomes prominent only
after the recent development of symplectic algorithms [Fen85] . In crucial aspects, C is
superior.
Remark 3.1. The above log fCs is not others but just formal vector fields for centered
Euler scheme or present called backward error analysis
s2 1
f¯ = f + (f f f − f (f, f )).
12 2
ż = f (z), z ∈ Rn . (4.1)
B-series methods: B-series were introduced by Harier and Wanner[HW74] . The Taylor
series of exact solution of (4.1) with initial value z(0) = z can be written as
h2 h3
z(h) = z + hf (z) + f (z)f (z) + f (f (z), f (z)) + f (z)f (z)f (z) + · · · .
2! 3!
(4.2)
B-series methods are numerical integrators zn+1 = Φh (zn ) whose Taylor series have
the same structure with real coefficients a(τ ):
a( )
Φh (z) = z + ha( )f (z) + h2 a( )f (z)f (z) + h3 f (f (z), f (z))
2!
+a( )f (z)f (z)f (z) + · · · , (4.3)
where coefficients a(τ ) are defined for all rooted trees and characterize the integrator.
Every numerical integrator (including R–K method) can be expanded into a B-
series as introduced and studied in[HW74] .
Definition 4.1 (rooted tree and forest). The set of rooted tree T and forest F are
defined recursively by
1◦ The tree , only one vertex belong to T ;
2◦ If τ1 , · · · , τn are n tree of τ , the forest u = τ1 , · · · , τn is the commutative
product of τ1 , · · · , τn ;
3◦ If u is a forest of F, then u = |τ | is a tree of T .
Let T = { , , , · · ·} be the set of rooted trees and let ∅ be the empty tree. For
τ1 , · · · , τn ∈ T , we denote by τ = [τ1 , · · · , τn ] the tree obtained by grafting the roots
of τ1 , · · · , τn to a new vertex which becomes the root of τ . Elementary differentials
Ff (τ ) are defined by induction as
Ff ( )(z) = f (z), Ff (τ )(z) = f (m) (z) Ff (τ1 ), · · · , Ff (τm )(z) . (4.4)
For real coefficients a(∅) and a(τ ), τ ∈ T a B-series is a series of the form
418 9. Formal Power Series and B-Series
h|τ |
B(f, a, z) = a(∅)Id + a(τ ) Ff (τ )(z) (4.5)
σ(τ )
τ ∈T
where Id stands for the identity; Id(z) = z and the scalars a(τ ) are the known nor-
malization coefficients[BSS96] . Now, we give following examples:
Example 4.2. The Taylor series of the exact solution of (4.1) can be written as a B-
1
series z(h) = B(f, e)(z0 ) with coefficients a(τ ) = e(τ ) = ,∀τ ∈ T.
γ(τ )
Example 4.3. The coefficient B-series for explicit Euler scheme a(τ ) = 0, ∀ τ ∈ T
except a( ) = 1.
Example 4.4. The coefficient B-series for implicit Euler scheme a(τ ) = 1, ∀ τ ∈ T .
|τ |−1
1
Example 4.5. The coefficient B-series for centered Euler scheme a(τ ) = ,
2
∀τ ∈ T.
Example 4.6. The coefficient B-series for trapezoidal scheme a( ) = 1, a( ) =
1 1 1
, a( ) = , a( ) = , · · · .
2 2 4
Example 4.7. The coefficient B-series for R–K method (A, b, c), a(τ ) = bT φ(τ ), ∀ τ ∈
T.
Partitions and skeletons: A partition pτ of a tree τ is obtained by cutting some of the
edges [CHV07] . The resulting list of trees is denoted by P (pτ ). Eventually, the set of all
partitions pτ of τ is denoted by P (pτ ). Now, given a partition pτ , the corresponding
skeleton χ(pτ ), as introduced in [CHV07] , is the tree obtained by contracting each tree of
P (pτ ) to a single vertex and by re-establishing the cut edges (see Tables 4.1 – 4.25).
We observe that a tree τ ∈ T has exactly 2|τ |−1 partitions pτ and that different parti-
tions may lead to the same P (pτ ). An admissible partition is a partition with at most
one cut along any part from the root to any terminal vertex. We denote AP τ as the set
of admissible partition of τ and by convention, we suppose that ∅ ∈ AP τ . We denote
#(pτ ) as number of subtrees. We denote this distinguished tree by R(pτ )(or rp ). We
denote P ∗ (pτ ) = P (pτ ) \ R(pτ ) as the list of forest that do not contain the root of τ .
We distinguish rp as the tree vp (or P (pτ )) whose root coincides with the root of τ .
This tree is usually referred to as a subtree of τ and we denoted by vp∗ (or P ∗ (pτ )) the
forest obtained by removing rp from vp . The above definition can be seen in Tables
4.1 – 4.25.
Theorem 4.8. Let a, b : T ∪{∅} → R be two mappings, with a(∅) = 1. Then B-series
B(f, a)(z) inserted into B(f, b)(·) is still a B-series
B(f, b) B(f, a)(z) = B(f, a · b)(z), (4.7)
(a · b) = b(∅) = b(∅), ∀ τ ∈ T, (a · b) = b(rp )a(vp∗ ), (4.8)
p∈AP(τ )
;
n
∀u = τ1 · · · τn ∈ F, a(u) = a(τi ). (4.9)
i=1
Table 4.1. The partitions of a tree of order 2 with associated skeleton and forest
• ·•
·
•·
τ
p •
•
•
χ(pτ ) •
' /
•
τ ••
P (p ) •
•
•
R(pτ ) •
∗ τ
∅ •
P (p )
#(pτ ) 1 2
pτ ∈ AP τ yes yes
420 9. Formal Power Series and B-Series
Table 4.2. The partitions of a tree of order 3 with associated skeleton and forest
• • •· • • ·• •· ·•
· · · ·
pτ • ·• •· ·•·
• • • •
•
χ(pτ ) • • •
'• •/ • •
P (p ) τ
• •• •• •••
• • • •
τ •
R(p ) • • •
∗ τ
∅ • • ••
P (p )
#(pτ ) 1 2 2 3
Table 4.3. The partitions of a tree of order 3 with associated skeleton and forest
• • • •·
·· ··
• • •· •
·· ··
pτ • •· • •·
•
• • •
τ •
χ(p ) • • •
'• /
• • •
• • •••
P (pτ ) • • •
•
• •
• •
R(pτ ) • •
' /
•
∗ τ
∅ • ••
P (p ) •
#(pτ ) 1 2 2 3
p ∈ AP
τ τ
yes yes yes no
9.4 Related B-Series Works 421
Table 4.4. The partitions of a tree of order 4 with associated skeleton and forest
• • • • • • • •
·· ·· ·· ··
• • •· • • ·• •· ·• · •· ·• •· ··•
·· ·• •
· ·
• ·• · · ·
· ·
pτ • • •· • ·•· •· ·• ·•·
• •
• • • • • • • • • •
τ •
χ(p ) • • • • • • •
' • / ' • / ' /' /' / ' / ' /
• •
• • • • • • • ••
•
••
•
••
•
••••
P (pτ ) • • • • • •
• •
• •
• • • • • • • •
τ • •
R(p ) • • • • • •
• •
∗
∅ • • • •• •• •••
P (p ) τ
• •
#(pτ ) 1 2 2 2 3 3 3 4
p ∈ AP τ
τ
yes yes yes yes yes no yes no
Table 4.5. The partitions of a tree of order 4 with associated skeleton and forest
• • • • • • •· • •· ·• • • •· • •· ·•
·· ·· ·· ·· ·· ·· ·· ··
• •· •· • • •·· •· •·
pτ · · · ·
• •· • • • •· •· •·
• • • •
• • • • • • • • •
χ(pτ ) • • • • • • •
'• /' • / ' • / '
•/ ' / ' / ' /
• • • • •
τ
• • • • • •• •• •• ••••
P (p ) • • • • • • • •
• • • •
• • • • • • •
•
R(pτ ) • • •
•
' / ' /' /
• • • •
∅ • • • , • •, •, •,•, •
P ∗ (pτ ) • • •
#(pτ ) 1 2 2 2 3 3 3 4
Table 4.6. The partitions of a tree of order 4 with associated skeleton and forest
• • •· • •· •· • •·
·· ·· ·· ··
• • • • • • • •
·· ·· ·· · ··
• •· • •· •· •·· •·· •·
pτ • ··• • • • ·• ··• ··•
•
• • •
•
• • • • • • •
•
χ(p ) τ
• • • • • • •
•
' / ' • / ' • / ' /' /' /' /
• • • • • •
• • • • • •• •• •• ••••
P (pτ ) • • • • • • • •
• •
•
• • • •
• • • •
R(pτ ) • • • •
' • /
• • • •
∅ • •• • • •••
P ∗ (pτ ) • • • •
#(pτ ) 1 2 2 2 3 3 3 4
p ∈ AP τ
τ
yes yes yes yes no no no no
Table 4.7. The partitions of a tree of order 5 with associated skeleton and forest
• • • •· • • • •· • • • ·• •· •· • •· • ·• • •· ·• •· •· ·•
·· ·· · ·· ·· ·· · · ·· · ·· ·· · ·
•· •·
τ
p • • • • • •
• • • • • • • • • • • •
τ •
χ(p ) • • • • • • •
• • • • • • • • • • • • •
R(p ) τ
• • • • • • •
#(pτ ) 1 2 2 2 3 3 3 4
Table 4.8. The partitions of a tree of order 5 with associated skeleton and forest
• • • • • • • •
·· ·· ··
• • •· •· • •·· • •·
·· ··
• • • • • • • •
·· ··
• •· • • •· • •·· •·
pτ
• ··• • • • • ··• ··•
• • •
• • • •
• • • •
χ(pτ ) • • • • • • •
•
'• /' • /' • /'
• • • •/' • • /' • /' • /' •/
• • • • • • • • • •• • •• • •• •
• • • • • •
P (p ) τ
• • •
•
•
•
• • •
•
• • • • • •
• •
R(p ) τ
• • • • •
•
' / ' •/ ' •/ ' •/
• •
∅ • • • •• •• ••
P ∗ (pτ ) •
• • • •
#(pτ ) 1 2 2 2 2 3 3 3
p ∈ AP τ
τ
yes yes yes yes yes no no no
Table 4.9. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
• • • • • • • •
·· ·· ·· ·· ··
•· •· •· •·· •· •·· •· •··
·· ·· ·· ·· ·· ··
• • • • • • • •
·· ·· ·· · ·· ·· ···
•· •· •· • · •· •· •·· •·
pτ ··• • • • ··• ··• ··• ··•
•
• • • • •
• • • • • • •
•
• • • • • • • •
τ
χ(p ) • • • • • • • •
'• • /'• • /'• • /' • /' •/ ' •/ ' • / ' /
• • • ••• ••• ••• ••• •• •• •
•• •• •• • • • •
P (pτ )
• • •
• • • • •
R(p ) τ • • •
'• • /' • /' • / ' •/ ' •/ ' • /
• • ••• •• •• •• ••••
P ∗ (pτ ) •• • • • • •
#(pτ ) 3 3 3 4 4 4 4 5
p ∈ AP τ
τ
no no no no no no no no
424 9. Formal Power Series and B-Series
Table 4.10. The partitions of a tree of order 5 with associated skeleton and forest
• • • • • • • • • • • • • • • •
• • •· • • ·· • • ·· • • ·• •· ···· • • ··· ··· • • ·····•
·· · · ·
•· ·• •·
τ
p • • • • •
• • • • • • • • • •
•
χ(p ) τ
• • • • • • •
• • • • • • • • • • • •
• • • • • • • • • •
•• •• ••
P (p ) τ
• • • • • • • • • • • •
• • • • • • • • • • • •
• • • • • • • • • •
R(p ) τ
• • • • • • • •
∅ • • • • •• •• ••
P ∗ (pτ )
#(pτ ) 1 2 2 2 2 3 3 3
Table 4.11. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
• • • • • • • • • • • • • • • •
• ·· ·• •· ·· • •· • •· ·· ·· • • ·· ·· ·• •· ·· ·• •· ·· ·• •· ·· ·· ·•
τ · ·· ·· · · · ··· · ·· · · ··· · · ··· · ·· ·· · ·· ···
p • • • • • • • •
• • • • • • • • • • • • • • • • • • •
• •
•
τ
p • • • • • • • •
• • • • • • • • • •
•
R(pτ ) • • • • • • •
•• •• •• ••• ••• ••• ••• ••••
P ∗ (pτ )
#(pτ ) 3 3 3 4 4 4 4 5
Table 4.12. The partitions of a tree of order 5 with associated skeleton and forest
• • • • • • • • · • • •·· • • • • • • • • •·· • • • · •
· · ·· ·· · ·
• •· •· • ·• ·•
·· •· •· ·
·· · ·
pτ • • • • • •· •· •·
• • •
• • • • • •
• •
χ(pτ ) • • • • • • •
'• • •/ '• •
/'• •
/'• •
/
• • • • • • •
• • • • • •
• • • •• •• ••
P (pτ ) • • • • • • • • •
• • • • • • • • •
• • • •
• • • •
R(pτ ) • • • •
' • / '• •
/'
• •
/'
• •
/
• •
∅ • • • • • •
P ∗ (pτ ) • • • •
3 1 2 2 2 2 3 3 4
p ∈ AP τ
τ
yes yes yes yes yes no no no
Table 4.13. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
• • • •
• • • • • • •
•
R(pτ ) • • •
' • /' • /' • /
• • • • • • ••• •• •• •• ••••
• • •
P ∗ (pτ )
#(pτ ) 3 3 3 4 4 4 4 5
pτ ∈ AP τ yes yes yes yes no no no no
426 9. Formal Power Series and B-Series
Table 4.14. The partitions of a tree of order 5 with associated skeleton and forest
• • • • • • • •
·· ·· ··
• •· •· • • •·· • •·
• • • • • ··• • • •· • • ··• •· ·• • ·•
·· ·· ·· ·· ·
pτ • • • •· • • • •·
• •
• • • • • • • •
•
χ(pτ ) • • • •
• • •
' • ' / ' / ' • • ' / ' • ' /
/ • / ' / /
• •• • • • • • • • • •
• • •• • • • •• • •• • • ••
• •• ••
P (pτ ) • •
• •
• • • • •
• • • • • • • • • •
• • • •
R(pτ ) • • •
Table 4.15. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
• • • • • • • •
·· ·· ·· ·· ··
•· •· •· •·· •·· •· •· •··
•· • • ···• •· ··• • ···• •· ··• •· ·• •· ···• •· ···•
·· · ·· · ·· ·· ·· ·· ·· ·· ··
pτ • •· • •· • • • •
• • •
• • • • • • • • •
• • • • • • • • •
• ··
χ(pτ ) • • • • • • • •·
' • / ' • • /' • • /' /' /' /' /
•• • • • • • • •••••
• • • • • • • • • • • • •
•• •• • • • •
P (pτ ) •
•
• • • • • •
• •
R(pτ ) • • • • •
Table 4.16. The partitions of a tree of order 5 with associated skeleton and forest
• • • •· •· • • • • • •· •· • • • •·
· · · · ·
• • • •· •· • • • •· • •· •· •· ·• • •·
·· ·· ·· ·· ·· ·
pτ • • • •· • • • •
•
• • • • • • • • •
•
χ(pτ ) • • • • • • •
Table 4.17. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
•· • • •· •· • •· • • •· •· •· •· •· •· •·
· · · · · · · · · · ·
•·· • •· •· •· • •·· ·• •· ·•· •· •· •·· •· •·· ·•·
·· ·· ·· ·· ·· ·· ·· ·· · ·· ·· ··
pτ • • •· • • • • •
• • • • • • •
• • • • • • • • • • • • • • •
χ(pτ ) • • • • • • • •
' • /'
• • / ' • • / ' • / ' • /' • / ' • /
•• • • • ••• ••• ••• ••• •••••
•• •• • • • •
P (pτ ) •
•
• • • •
• • • •
τ • • • •
R(p ) •
' / /
' • / ' • / • ' •
• • ••• • • • ••• • • • • • • •
• • • • • •
P ∗ (pτ )
#(pτ ) 3 3 3 4 4 4 4 5
p ∈ AP τ
τ
no yes yes no no no no no
428 9. Formal Power Series and B-Series
Table 4.18. The partitions of a tree of order 5 with associated skeleton and forest
• •· • • • •· • •
• • ·· • • • • ·· • •
• • • • •· • • · • • · ·• • • •· · • •· •
·· ·· · ·· ·· ·· ·· ·· · · ·
pτ • • • • • •· • •
•
• • • • • • • • •
•
χ(pτ ) • • • • • • •
#(pτ ) 1 2 2 2 2 3 3 3
pτ ∈ AP τ yes yes yes yes yes no yes yes
Table 4.19. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
•· •· • • •· •· •· •·
· · · · · ·
• •· ·• •· • ·• • •· ·• •· •· ·• • •· ··• •· • ··• •· •· ·• •· •· ··•
·· ·· ·· · ·· ·· ··· ·· · ·· · · ·
·· · ·
·· · · ·
pτ • • •· • •· • • •
• • •
• • • • • • • • • • • •· • • • • • • •
·
χ(pτ ) • • • • • ·• • •
' /' /' ' /' /' /' /
• • • • •• / • • • •
•• •• • • • • • • • • • • • • • • • • • •
P (pτ ) • • •• • • • •
• • • • • • • •
• •
R(pτ ) • • • • • •
' • / ' • /
• • • • • •• ••• ••• ••• ••••
P ∗ (pτ ) • •
#(pτ ) 3 3 3 4 4 4 4 5
pτ ∈ AP τ yes yes yes yes no no yes no
9.4 Related B-Series Works 429
Table 4.20. The partitions of a tree of order 5 with associated skeleton and forest
• • • • •· • • • • • •· ·• • • •· •
·· · · · ·
• • • •· • ·• • ·• •· • • ·•· •· ·• •· ·•
·· ·· ·· ·· ··
pτ • • • • • • • •
• • • • • • • • • •
•
χ(pτ ) • • • • • • •
• • • • • • •
• • • • • • • • • •
• •
R(pτ ) • • • • • •
•
' / ' • /
•
• •
∅ • • • • • • • •
• •
P ∗ (pτ )
#(pτ ) 1 2 2 2 2 3 3 3
p ∈ AP τ
τ
yes yes yes yes yes yes yes yes
Table 4.21. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
•· • • ·• • ·• •· ·• • ·• •· • •· ·• •· ·•
·· ·· • •·· • ··•·· • •·· • ··• ·· ·· ·· ··
• ·• • ·• · · · · • • • •
· · ·· · ··· ·· ··· ·· ··· ··· ···
pτ •· •· • •· • • • •
• • • • • • • •
• • • • • • • • • • • •
• •
χ(pτ ) • • • • • • • •
' • /
•• ••
•• •
• • • •
• • ••• ••• ••• ••• •••••
P (pτ ) •• •• • • • • •
•
• • • • •
• • •
R(pτ ) • • • • •
• • • •
• • • • ••• •• •• ••• ••••
∗
P (p ) τ • • • •
τ
#(p ) 3 3 3 4 4 4 4 5
p ∈ AP
τ τ
no no yes no no no no no
430 9. Formal Power Series and B-Series
Table 4.22. The partitions of a tree of order 5 with associated skeleton and forest
• • • • • • •· • • • • • •· ·• •· •
·· ·· ·· ·· ··
• • •· • •· •· • •·
·· ·· ··
• • • • • • • •
pτ ··· · ··
• • • • • • • •
• •
• • • • • • • •
•
χ(pτ ) • • • • • • •
• • • •
' • / ' • •/ ' /' • /' • /' /' • / ' /
• • • • • ••
• • • • • • • •• •• • •
• •• • ••
P (pτ ) • • • •
• • • •
• • • •
• • • • • • • •
R(pτ ) • • • • •
•
' • •/ ' / ' / ' • /
• • • • •
∅ • • • •• •
• •
P ∗ (pτ ) • •
#(pτ ) 1 2 2 2 2 3 3 3
pτ ∈ AP τ yes yes yes yes yes no yes no
Table 4.23. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
•· • • ·• • ·• •· ·• •· ·• •· • • ·• •· ·•
·· · · ·· ·· ·· ·· ·· · ·· ··
• •·· •· •· • •· •·· •·
·· ·· ·· ·· ··
• • • • • • • •
pτ ·· ·· ·· ·· ·· ··
•· • •· • •· •· •· •·
• • • •
• • • • • • • • • •
• • • • • • • •
τ
χ(p ) • • • • • • • •
' •/ ' •/
•• • • • •
• • •
•• • •• •• ••• ••• ••• ••• •••••
P (pτ ) • • • • •
• •
• • • • • •
R(pτ ) • •
' •/ • ' •/ • • •
• • • • • ••• •• •• •• ••• •
• • • •
P ∗ (pτ ) • •
#(pτ ) 3 3 3 4 4 4 4 5
p ∈ AP τ
τ
no no no no no no no no
9.4 Related B-Series Works 431
Table 4.24. The partitions of a tree of order 5 with associated skeleton and forest
• • • •· • •· • •
·· ··
• • • • •· • • • • • • • •· ·• •· •
·· ·· ·· ·· ·· ·
• •· • • •· •· • •·
· ·
pτ • •· • • • • • •·
• •
• • • • • • • •
•
χ(pτ ) • • • • • • •
• •
' / ' • /' /
• • • ' • • / '• • / ' • / • ' • /
•• • • • •
• • • • •• • ••
•
•• •
P (pτ ) • • • • • • •
• •
• • • • • • •
• • • • • •
• •
R(p ) τ
• • • • •
•
•
' / • ' • /
• • • • •
∅ • • • • •
• •
P ∗ (pτ ) • •
#(pτ ) 1 2 2 2 2 3 3 3
pτ ∈ AP τ yes yes yes yes yes no yes no
Table 4.25. Continuous partitions of the above tree of order 5 with associated skeleton and
forest
• •· •· •· •· •· • •·
·· · · ·· · ·
• • • • •· ·• •· ··• • • •· ·• •· ·• •· ··•
·· ·· ·· ·· ·· ·· ·· ·· ·· ··
•· · •· • • •· · •· •· •·
· · · · · ·
pτ •· •· • • •· •· •· •·
• •
• • • • • • •
• • •
• • • • • • • • • •
χ(pτ ) • • • • • • •
•
•/
• /' ' •
• • • • • •
•• •••
P (pτ ) • •• • ••
• • • • • • • • • • • •••••
• • •
•
• •
• • • • • •
R(pτ ) • •
• • • • •
• •
• •• ••• •• •• •• •• • •
•• • • • •
P ∗ (pτ )
#(pτ ) 3 3 3 4 4 4 4 5
pτ ∈ AP τ no no yes no no no no no
432 9. Formal Power Series and B-Series
Remark 4.10. The composition law for the trees of order ≤ 5 is listed in Example
4.22.
Remark 4.11. The Substitution law for the (backward error) trees of order ≤ 5 is
listed in Example 4.24.
Remark 4.12. The Substitution law for the trees of order ≤ 5 is listed in Example
4.23.
Modified integrators (called generating function method or preprocessed vector field
integrators): Let Ψf,h is the exact h-flow for Equation (4.1) which is a B-series with
1
coefficient e(τ ) = . Consequently, the coefficient b̆(τ ) of the modified differen-
γ(τ )
tial equation for Φf,h = B(f, a) is obtained from
Backward error analysis (called formal vector field, modified equation or postpro-
cessed vector field): The modified differential equation of a method Ψf,h = B(f, e)
is obtained by putting Φf,h equal to the exact flow. Its coefficient b(τ ) is therefore
obtained from
(b ∗ e)(τ ) = a(τ ), ∀ τ ∈ T. (4.14)
1
Remark 4.13. Substituting the expression given in (4.13) into (4.14) gives b̆∗b∗a =
γ
a. Therefore, b̆ and b(τ ) are inverse elements for substitution law ∗
Proposition 4.14. Using formulae (4.13) and (4.11) in Example 4.23, we easily ob-
tain modified centered Euler scheme of sixth order first find in[CHV07] :
h2 @ (2) A h4 3 (4)
ż = f (z) + f (f, f ) − 2f (1) f (1) f + f (f, f, f, f )
24 120 48
1 (3) 1 (2)
− f (f, f, f f ) + f (f, f (f, f )) − f (2) (f, f (1) f (1) f )
(1) (2)
4 4
2 (2) (1) 3 1
+ f (f f, f f ) − f (1) f (3) (f, f, f ) + f (1) f (2) (f, f (1) f )
(1)
4 12 2
1 (1) (1) (2)
− f f f (f, f ) + f f f f f + O(h6 ).
(1) (1) (1) (1)
4
Proof. First, we must point out b̆(τ ) = 0, ∀ |τ | = even. We calculate coefficient b̆( )
as follows
Table 4.26. Coefficients σ(τ ), γ(τ ), b̆(τ ), and b(τ ) for trees of order 5
τ ∅
σ(τ ) 1 1 2 1 6 1 2 1
γ(τ ) 1 2 3 6 4 8 12 24
σ(τ ) 24 2 2 6 1 1 2 1 2
γ(τ ) 5 10 20 20 40 30 60 120 15
b̆(τ ) 1/80 −1/240 1/120 −1/80 1/240 −1/120 −1/240 1/120 1/240
b(τ ) 7/240 1/240 1/80 −7/240 −1/240 −1/80 1/240 1/80 −1/240
Properties of logarithmic map has been discussed in Proposition 2.1. Using formula
of Example 4.24, determined ω(τ )(= b(τ )) recursively , because a(τ ) = 0 ∀τ ∈ T
except a( ) = 1.
For example: from 14 formula of Example 4.24, we have
11 1 1 1 1 1 1 1 1 1 1 1 1
b( ) + (− ) + ( )(− )( ) + (− )( ) + (− ) + ( ) + 2( )( )
122 2 2 6 2 2 3 2 4 6 6 6 4
1 1 1 1 1 1 1 1 1
+( )( ) + 2( )( ) + ( )(− ) + 3( )(− ) + = 0,
3 4 3 3 24 2 8 2 30
then we get
1
ω( ) = b( )= .
20
τ ∅
ω(τ ) –1/30 –1/60 1/30 1/30 1/10 1/20 3/20 1/5 1/60
Definition 4.19. (Lie derivative of B-series) Let b(τ ) with b(∅) = 0 and a(τ ) be the
coefficient of two B-series and let z(t) be a formal solution of the differential equation
hż(t) = B(b, z(t)). The Lie derivatives of the function B(a, z(t)) with respect to the
vector field B(b, z(t)) is again B-series
d
h B(a, z(t)) = B(∂b a, z(t)). (4.18)
dt
Exercise 4.20. [HLW02] Prove that the coefficient of modified differential equation are
recursively defined by b(∅) = 0, b(·) = 1 and
|τ |
1
b(τ ) = a(τ ) − ∂bj−1 b(τ ), (4.20)
j!
j=2
1
Proposition 4.21. The above-mentioned formula (4.20) is just formula b ∗ =
γ(τ )
a(τ ) namely
|τ |
1 1
∂bj−1 b(τ ) = b(τ ) ∗ . (4.21)
j! γ(τ )
j=1
Proof. Note that formula (4.23) in Example 4.24 and Tables 4.1 – 4.25, can obtain
this results directly.
436 9. Formal Power Series and B-Series
Example 4.22. The composition laws for the trees of order ≤ 5 are
a · b( ) = b(∅) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( )2 + 2b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( ) + b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( )3 + 3b( ) · a( )2 + 3b( ) · a( )
+b( )
a · b( ) = b(∅) · a( ) + b( ) · a( )a( ) + b( ) · a( ) + b( ) · a( )2
+b( ) · a( ) + b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( ) + b( ) · a( )2 + 2b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( ) + b( ) · a( ) + b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( )4 + 4b( ) · a( )3 + 6b( ) · a( )2
+4b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( )3 + b( ) · a( ) · a( )2
+2b( ) · a( ) · a( ) + b( ) · a( ) + b( ) · a( )2
+2b( ) · a( )2 + 2b( ) · a( ) + b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + 2b( ) · a( ) · a( ) + b( ) · a( ) + b( ) · a( )2
2
+2b( ) · a( ) + 2b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( )3 + 3b( ) · a( )2 + 3b( ) · a( )
+b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( ) · a( ) + b( ) · a( )2 + b( ) · a( )
+b( ) · a( ) + b( ) · a( ) + b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( ) · a( ) + b( ) · a( ) · a( ) + b( ) · a( )
+b( ) · a( ) + b( ) · a( )2 + b( ) · a( ) + b( ) · a( ) + b( )
9.4 Related B-Series Works 437
a · b( ) = b(∅) · a( ) + b( ) · a( ) + b( ) · a( ) + b( ) · a( )2
+2b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + b( ) · a( ) + b( ) · a( ) + b( ) · a( )
+b( ) · a( ) + b( )
a · b( ) = b(∅) · a( ) + 2b( ) · a( )2 + b( ) · a( ) · a( ) + b( ) · a( )2
+b( ) · a( ) + 2b( ) · a( ) + b( ) · a( ) + b( )
Example 4.23. The substitution law ∗ defined in for the trees of order ≤ 5.
b ∗ a( ) = a( )b( )
b ∗ a( ) = a( )b( ) + a( )b( )2
b ∗ a( ) = a( )b( ) + 2a( )b( )b( ) + a( )b( )3
b ∗ a( ) = a( )b( ) + 2a( )b( )b( ) + a( )b( )3
b ∗ a( ) = a( )b( ) + 3a( )b( )b( ) + 3a( )b( )2 b( ) + a( )b( )4
b ∗ a( ) = a( )b( ) + a( )b( )b( ) + a( )b( )2 + a( )b( )b( )
+2a( )b( )2 b( ) + a( )b( )2 b( ) + a( )b( )4
b ∗ a( ) = a( )b( ) + a( )b( )b( ) + 2a( )b( )a( ) + a( )b( )2 b( )
+2a( )b( )2 b( ) + a( )b( )4
+a( )b( )4
b ∗ a( ) = a( )b( ) + 4a( )b( )b( ) + 6a( )b( )2 b( )
+4a( )b( )3 b( ) + a( )b( )5
b ∗ a( ) = a( )b( ) + a( )b( )b( ) + 2a( )b( )b( ) + a( )b( )b( )
+a( )b( )2 b( ) + 2a( )b( )2 b( ) + a( )b( )2 b( )
+2a( )b( )b( )2 + 2a( )b( )3 b( ) + 2a( )b( )3 b( )
+a( )b( )5
b ∗ a( ) = a( )b( ) + 2a( )b( )b( ) + 2a( )b( )b( )
+a( )b( ) b( ) + 2a( )b( )2 b( ) + 3a( )b( )b( )2
2
+a( )b( )5
+a( )b( )2 b( )
+a( )b( )5
Example 4.24. The substitution law ∗ defined in for the trees of order ≤ 5
b ∗ e( ) = e( )b( )
b ∗ e( ) = e( )b( ) + e( )b( )2
+e( )b( )4
(4.23)
+e( )b( )5
+e( )b( )5
[Bak05] H. F. Baker: Alternants and continuous groups. Proc. London Math. Soc., 3:24–47,
(1905).
[Bro00] Ch. Brouder: Runge–Kutta methods and renormalization. Euro. Phys. J. C, 12:521–
534, (2000).
[BSS96] J. C. Butcher and J. M. Sanz-Serna: The number of conditions for a Runge–Kutta
method to have effective order p. Appl. Numer. Math., 22:103–111, (1996).
[CEFM08] D. Calaque, K. Ebrahimi-Fard, and D. Manchon: Two Hopf algebra of trees inter-
acting. arXiv: 0806.2238 v 2, (2008).
[CHV05] P. Chartier, E. Hairer, and G. Vilmart: A substitution law for B-series vector fields.
Technical Report 5498, INRIA, (2005).
[CHV07] P. Chartier, E. Hairer, and G. Vilmart: Numerical integration based on modified
differential equations. Math. Comp., 76(260):1941–1953, (2007).
[CHV08] P. Chartier, E. Hairer, and G Vilmart: Composing and substituting S-series and B-
series of integrators and vector fields. Preprint, www.irisa.fr/ipso/fichiers/algebraic.pdf,
(2008).
[CK98] A. Connes and D. Kreimer: Hopf algebra, renormazation and noncommutative geom-
etry. Communications in Mathematical Physics, 199:203–242, (1998).
[Dyn46] E. B. Dynkin: Normed Lie algebra and analytic groups, volume 1. Amer. Math. Soc.
(translation), (1946).
[Fen85] K. Feng: On difference schemes and symplectic geometry. In K. Feng, editor, Pro-
ceedings of the 1984 Beijing Symposium on Differential Geometry and Differential Equa-
tions, pages 42–58. Science Press, Beijing, (1985).
[Fen92] K. Feng: Formal power series and numerical methods for differential equations. In
T. Chan and Z. C. Shi, editors, International conf. on scientific computation, pages 28–35.
World Scientific, Singapore, (1992).
[Fen93a] K. Feng: Formal dynamical systems and numerical algorithms. In K. Feng and Z.
C. Shi, editors, International conf. on computation of differential equations and dynamical
systems, pages 1–10. World Scientific, Singapore, (1993).
[Fen93b] K. Feng: Symplectic, contact and volume preserving algorithms. In Z.C. Shi and
T. Ushijima, editors, Proc.1st China-Japan conf. on computation of differential equations
and dynamical systems, pages 1–28. World Scientific, Singapore, (1993).
[Hai94] E. Hairer: Backward analysis of numerical integrators and symplectic methods. Annals
of Numer. Math., 1:107–132, (1994).
[Hau06] F. Hausdorff: Die symbolische exponentialformel in der gruppentheorie. Berichte der
Sachsischen Akad. der Wissensch., 58:19–48, (1906).
[HLW02] E. Hairer, Ch. Lubich, and G. Wanner: Geometric Numerical Integration. Num-
ber 31 in Springer Series in Computational Mathematics. Springer-Verlag, Berlin, (2002).
[HW74] E. Hairer and G. Wanner: On the Butcher group and general multivalue methods.
Computing, 13:1–15, (1974).
[Lie88] S. Lie; Zur theorie der transformationsgruppen. Christiania, Gesammelte Abh., Christ.
Forh. Aar., 13, (1988).
442 Bibliography
[Mur06] A. Murua: The Hopf algebra of rooted trees, free Lie algebra, and Lie series. Foun-
dations of Computational Mathematics, 6(4):387–426, (2006).
[Olv93] P. J. Olver: Applications of Lie Groups to Differential Equations. GTM 107. Springer-
Verlag, Berlin, Second edition, (1993).
[Ote91] J. A. Oteo: The Baker–Campbell–Hausdorff formula and nested commutator identi-
ties. J. of Math. Phys., 32(2):419–424, (1991).
[OW00] B. Owren and B. Welfert: The Newton iteration on Lie groups. BIT, 40(1):121–145,
(2000).
[Owr06] B. Owren: Order conditions for commutator-free Lie group methods. J. Phys. A:
Math. Gen., 39:5585–5599, (2006).
[Rei99] S. Reich: Backward error analysis for numerical integrators. SIAM J. Numer. Anal.,
36:475–491, (1999).
[SS96] J. M. Sanz-Serna: Backward Error Analysis for Symplectic Integrators. In J. E. Mard-
sen, G. W. Patrick, and W. F. Shadwick, editors, Integration Algorithms and Classical Me-
chanics, pages 193–206. American Mathematical Society, New York, (1996).
[SS97] J. M. Sanz-Serna: Geometric integration. In The State of the Art in Numerical Analysis
(York, 1996), volume 63 of Inst. Math. Appl. Conf. Ser. New Ser., pages 121–143, Oxford
Univ. Press, New York, (1997).
Chapter 10.
Volume-Preserving Methods for Source-Free
Systems
here etf denotes the flow of system (1.1) and (etf (x))∗ the Jacobian of etf at x. Thus,
volume-preserving
schemes
are required for computing the numerical solution of
∂ xn+1
(1.1). If det = 1, we call this scheme volume-preserving, where xn denotes
∂ xn
the numerical solution at step n.
We know that the phase flow of Hamiltonian system preserves phase volume in-
variable. The source-free system is more general than the Hamiltonian system, we
must prove that the phase flow preserving phase volume is invariable, considering the
dynamic system (1.1), its phase flow is
for any t, using the formula for changing variables in a multiple integral gives
-
∂ gt x
V (t) = det d x.
D(0) ∂x
∂ gt x
Calculating by formula (1.2), we find
∂x
∂ gt x ∂f
=E+ t + O(t2 ), as t → 0.
∂x ∂x
but
det (E + At) = 1 + t tr A + O(t2 ), t → 0,
n
where tr A = aii . Therefore
i=1
-
V (t) = [1 + t div f + O(t2 )]d x, (1.4)
D(0)
-
d V (t)
= div f d x.
d t t=0 D(0)
d V (t)
and if div f = 0, = 0. This completes the proof.
dt
In particular, for Hamiltonian equation
∂ ∂H ∂ ∂H
div f = − + = 0,
∂p ∂q ∂q ∂p
Liouville’s theorem is proved specially.
xn+1 + xn
xn+1 = xn + τ f , (2.1)
2
where τ is the step size in t. We then have
∂ xn+1 xn+1 + xn 1 ∂ xn+1 1
= IN + τ Df + IN ,
∂ xn 2 2 ∂ xn 2
τ
∂ xn+1 IN + Df (x∗ )
= 2 .
∂ xn τ
IN − Df (x∗ )
2
∂f x + xn ∂ xn+1
Here, Df = fx = ≡ B = (bij ), x∗ = n+1 . The condition det =
∂x 2 ∂xn
τ
|IN + Df (x∗ )|
1, now requires 2
τ = 1. Let P (λ) = |Df (x∗ ) − λIn | be the character-
|IN − Df (x∗ )|
2
istic matrix of Df (x∗ ). Since
% &
% &
τ τ ∗ 2 2
IN + Df (x∗ ) Df (x ) + IN P
2 2 τ τ
= % & = (−1)
N
% &,
IN − τ Df (x∗ ) τ 2 P −
2
2 − Df (x∗ ) − IN τ
2 τ
Let us consider some particular cases of N to show that scheme (2.1) is not always
volume preserving.
Case 2.1. In this case, we have
N
∂fi
Since = 0, i.e., tr B = 0, then P (λ) = λ2 + b11 b22 − b12 b21 , and
i=1
∂xi
P (−λ) = P (λ).
Thus, the scheme (2.1) is always volume-preserving for source-free systems of dim.2.
Case 2.2. Here
P (λ) = −λ3 + (b11 + b22 + b33 )λ2 − cλ + |B| = −λ3 − cλ + |B|, (2.3)
where
b11 b12 b22 b23 b11 b13
c= + + .
b21 b22 b32 b33 b31 b33
446 10. Volume-Preserving Methods for Source-Free Systems
The volume-preserving condition for Euler method is now |B| = 0. For example,
(ABC flow) when system (1.1) takes the form
dx
= cy − bz,
dt
dy
= az − cx, a, b, c ∈ R,
dt
dz
= bx − ay,
dt
we have |B| = 0. For this dynamical system, centered Euler method is volume-
preserving.
Lemma 2.3. Let P (λ) be the characteristic polynomial of matrix AN ×N , then
P (λ) = |A − λIN | = (−1)N λN − P1 λN −1 + P2 λN −2 + · · · + (−1)N PN , (2.4)
where
N
P1 = aii = tr A,
i
N a
ii aij
P2 = ,
aji ajj
i<j
aii aij aik (2.5)
N
P3 = a ,
ji ajj ajk
i<j<k
aki akj akk
···
PN = |A|.
Using Lemma 2.3, we can discuss the case N = 4.
Case 2.4. At this time,
P (λ) = λ4 − P1 λ3 + P2 λ2 − P3 λ + |B|.
Since P1 = tr (B) = 0, then P (−λ) = (−1)4 P (λ) requires P3 = 0.
It must be pointed out that, when N increases, more increasing number of condi-
tions is required for system (2.1) to be volume-preserving, and it seems impossible to
satisfy all these condition. But fortunately, for the special case when system (1.1) is
Hamiltonian, i.e.,
5 6
O −Ik
f = J∇H, J = , N = 2k.
Ik O
From Lemma 2.5, we know that P (−λ) = (−λ)2k P (λ) is valid when system
(1.1) is Hamiltonian, so Euler method is volume-preserving for Hamiltonian systems.
In fact, the method is even symplectic for Hamiltonian systems, that is to say it also
preserve the symplectic structure of Hamiltonian systems which is a much stronger
property than volume-preserving.
···
⎧
⎪
⎪
d x1
= 0,
⎪
⎪
⎪
⎪
dt
⎪
⎪
⎪
⎨ ..
.
(2.9)
⎪
⎪ d xN −1
⎪
⎪ = 0,
⎪
⎪ dt
⎪
⎪
⎪
⎩ d xN = fN (x1 , · · · , xN −1 ).
dt
448 10. Volume-Preserving Methods for Source-Free Systems
The first order explicit Euler method can be applied to them to get the exact solutions
of them, i.e., the phase flows of them. Using the composition method[QZ92] , we can
construct first order explicit Euler volume-preserving scheme for system (2.6). The
adjoint of this scheme is obtained from the implicit Euler method and is also explicit.
Composing these two schemes, we get a reversible explicit. This process can be ex-
pressed by formal power series as shown below.
From Chapter 9, we know the flow of (1.1) can be represented by power of series.
∞
1 ∗k
eτf = 1N + τ k ek,f , ek,f : RN −→ RN , ek,f = f 1N ,
k!
k=1
N
∂
where f ∗ denotes the first order differential operator , f ∗ = fi , f ∗2 = f ∗ ×
i=1
∂xi
f ∗ , f ∗ 3 = f ∗ × f ∗ × f ∗ , · · · , 1N is the identity vector function, 1N (x) = x. For
simplicity, we just write out
eτA · eτB = eτcτ , (2.10)
the first several terms are
τ
cτ = A + B + [A, B] + o(τ 2 ),
2
where [A, B] = A∗ B − B∗ A is the Lie bracket of A and B, A∗ , B∗ denotes the
Jacobian matrix of A and B .
We now rewrite system of Equations (2.7) – (2.9) in compact form as
dx
= ai (x), ai = (0, · · · , 0, fi , 0, · · · , 0)T , i = 1, 2, · · · , N. (2.11)
dt
These integrable systems have flow
inf
eτai = 1N + τ k ek,ai , i = 1, 2, · · · , N. (2.12)
k=1
inf
inf
τk
eτai (x) = x + τ k ek,ai (x) = x + a∗i k 1N (x) = x + τ ai (x). (2.13)
k!
k=1 k=1
This means the concatenation eτaN × eτaN −1 × · · · × eτa1 approximates the flow eτf to
the first order of τ .
Because the equations in the system (2.11) are all source-free, their flows are all
volume-preserving and the concatenation of them remains volume-preserving, so
10.3 Source-Free System 449
det (eτaN × eτaN −1 × · · · × eτa1 )(x) ∗
= det eτaN (xN −1 ))∗ × det (eτaN −1 (xN −2 ))∗ × · · · × det (eτa1 (x0 ))∗ = 1,
of second order, but is still explicit. We can use theory of composition[QZ92] to con-
struct symplectic scheme of arbitrary order.
n
∂ ai (x)
div a(x) = = 0, ∀x ∈ Rn , (3.1)
i=1
∂ xi
through equations
dx
= ẋ = a(x), (3.2)
dt
here and hereafter, we use the coordinate description and matrix notation
to the solutions of systems and “proper” algorithms may generate remarkably right
ones.
But how does one evaluate a numerical algorithm to be proper for source-free
systems? It is well known that intrinsic to all source-free systems there is a volume
form of the phase space Rn , say
such that the evolution of dynamics preserves this form. In other words, the phase flow
eta , of source-free system (3.2), satisfies the volume-preserving condition
(eta )∗ α = α, (3.5)
or equivalently,
∂ eta (x)
det = 1, ∀ x ∈ Rn , t ∈ R. (3.6)
∂x
In addition to this, eta satisfies the group property in t,
In fact, (3.5) and (3.7) completely describe the properties of the most general
source-free dynamical systems. This fact suggests that a proper algorithmic approxi-
mation gas to phase flow esa for source-free vector field a : Rn → Rn should satisfy
these two requirements. However, the group property (3.7) is too stringent in general
for algorithmic approximations because only the phase flows satisfy it. Instead of it, a
weaker requirement, i.e.,
is reasonable and practicable for all vector fields a : Rn → Rn . We call such algo-
rithmic revertible approximations, that means gas always generate coincident forward
and backward orbits.
As for the volume-preserving property (3.5), it characterizes the geometric struc-
ture —volume-preserving structure—of source-free systems. Our aim here is just
to construct difference schemes preserving this structure, which we call volume-
preserving schemes, in sense that the algorithmic approximations to the phase flows
satisfy (3.5) for the most general source-free systems.
Lemma 4.2 says that there are no consistent analytic approximations to the expo-
nential function sending sl(n) into SL(n) at the same time other than the exponential
itself. This shows that it is impossible to construct volume-preserving algorithms ana-
lytically depending on source-free vector fields. Thus we have:
Theorem 4.3 (Feng-Shang). All the conventional methods including the well-known
Runge–Kutta methods, linear multistep methods and Euler methods (explicit, implicit
and centered) are non-volume-preserving.
The above lemma tell us we cannot construct volume-preserving scheme for all
source-free system. But we can split class sl(n) to subclass and perhaps in subclass,
there exists volume-preserving scheme.
In Subsection 10.2.1, we get some condition for centered Euler scheme to be
volume-preserving scheme. It is the best elucidation.
Consequently, to construct volume-preserving algorithms for source-free systems,
we must break through the conventional model and explore new ways.
n
α= ai (x)d xi . (5.5)
i=1
α = δβ. (5.6)
But for the 2-form β, there exists a skew symmetric tensor of order 2, b = (bik )1≤i.k≤n ,
bik = −bki , so that
n
β= bik d xi ∧ d xk . (5.7)
i,k=1
and from Equations (5.5) and (5.6), we get (5.4). The proof is completed.
N0 : bik = 0, |i − k| ≥ 2, (5.10)
Then, simple calculations show that all bk,k+1 are uniquely determined by quadra-
ture
- x2
b12 = a1 d x2 , (5.14)
-0 xk+1
∂b
bk,k+1 = ak + k−1,k d xk+1 , 2 ≤ k ≤ n − 2, (5.15)
∂ xk−1
-0 xn % & - xn−1
∂b
bn−1,n = an−1 + n−2,n−1 d xn − an |xn =0 d xn−1 . (5.16)
0 ∂ xn−2 0
n−1 % &T
∂ bk,k+1 ∂b
a= a(k) , a(k) = 0, · · · , 0, , − k,k+1 , 0, · · · , 0 , (5.17)
∂ xk+1 ∂ xk
k=1
or in components, ⎧
⎪
⎪
∂b
a1 = 12 ,
⎪
⎪
⎪
⎪
∂ x2
⎪
⎪ ∂b ∂b
⎪
⎪ a2 = − 12 + 23 ,
⎪
⎪ ∂ x1 ∂ x3
⎨
..
. (5.18)
⎪
⎪
⎪
⎪ ∂b ∂b
⎪
⎪ an−1 = − n−2,n−1 + n−1,n ,
⎪
⎪ ∂ x ∂ xn
⎪
⎪
n−2
⎪
⎪ ∂ bn−1,n
⎩ an = − .
∂ xn−1
m
a= a(i) , (6.1)
i=1
with smooth fields a(i) :Rn → Rn (i = 1, · · · , m). Suppose that, for each i =
1, · · · , m, Gτi is an approximation of order p to eτa(i) , the phase flow of the system
10.6 Construction of Volume-Preserving Schemes 455
1
associated to the field a(i) , in the sense that lim (Gτ (x) − eτa(i) (x)) = 0 for all
τ →0 τp i
x ∈ Rn with some p ≥ 1. Then, we have:
1◦ For any permutation (i1 i2 · · · im ) of (12 · · · m), the compositions
−1
τ
1 Gi1 i2 ···im := Gτim ◦ · · · ◦ Gτi2 ◦ Gτi1 , τ
1 Gi1 i2 ···im := −τ
1 Gi1 i2 ···im (6.2)
2(l+1) G
τ
= 2l Gαl τ ◦ 2l Gβl τ ◦ 2l Gαl τ , (6.7)
with 1
αl = (2 − 2 (2l+1) )−1 , βl = 1 − 2αl < 0, (6.8)
as a revertible approximation, of order 2(l + 1), to eτa .
where
∂a1 (x)
a1 (x) = a(x), a2 (x) = a(x),
∂x
(6.10)
∂ak−1 (x)
ak (x) = a(x), k = 1, 2, · · · .
∂x
This implies that 1 Gτ(12···m) is an approximation, of order one, to eτa , which provides
the proof needed.
In[QZ92] that 2 giτ1 i2 ···im and 2 giτ1 i2 ···im , defined by Equation (6.2), are revertible
approximations, of order 2, to eτa , the conclusion 1◦ of the Lemma 6.1 is proved.
2◦ By assumption, we have that for x ∈ Rn and τ ∼ 0,
1 2
Gτi (x) = x + τ a(i) (x) + τ 2 a(i) (x) + O(τ 3 ), i = 1, 2, · · · , m. (6.14)
2
Taylor expansion of the right hand side of Equation (6.4) with (i1 i2 · · · im ) =
(12 · · · m) yields
⎛ ⎞
m
1
m
τ
2 G(12···m) (x) =x+τ a(i) (x) + τ 2 ⎝ a(i) a(j) ⎠ (x) + O(τ 3 ), τ 0.
i=1
2 i,j=1
(6.15)
Here, we have used the convention
∂a(x)
(ab)(x) = (a∗ b)(x) = a∗ (x)b(x), a∗ (x) = , (6.16)
∂x
for a, b : Rn → Rn . However, we have
*m +
m
m
m
2 (i)
a = a∗ a = a a(j) = (a(i) )∗ a(j) = a(i) a(j) . (6.17)
i=1 ∗ j=1 i,j=1 i,j=1
So
1
eτa (x) = x + τ a(x) + τ 2 a2 (x) + O(τ 3 ) = 2 Gτ(12···m) (x) + O(τ 3 ), τ ∼ 0.
2
This shows that 2 Gτ(12···m) is an approximation, of order 2, to eτa . By direct verifica-
tion, this is revertible if each component Gτi is revertible.
The conclusion 3◦ directly follows from Qin-Zhu’s paper[QZ93] .
10.6 Construction of Volume-Preserving Schemes 457
2g
τ 2 ◦ 1G 2 ,
:= 1 G
τ τ
τ
2g
τ
2
= 1G 2 ◦ 1G
τ
2 G = Sn−1 ◦ · · · ◦ S2 ◦ S1 ◦ S1 ◦ S2 ◦ · · · ◦ Sn−1
τ 2 2 2 2 2 2
(6.20)
is a volume-preserving approximation, of order 2, to eτa .
3◦ If each Skτ is revertible, then the so-constructed 2 Gτ is revertible too.
4◦ From the above constructed revertible algorithmic approximation 2 g τ or
τ
2 G , we can further recursively construct revertible approximations, of all even
orders, to eτa according to the process of Lemma 6.1.
458 10. Volume-Preserving Methods for Source-Free Systems
Remark 6.4. If a has essentially Hamiltonian decompositions other than (5.17) and
(5.14) – (5.16), then one can construct volume-preserving difference schemes corre-
sponding to these decompositions in a similar way to the above.
⎧ 1
⎪ τ 1
⎪
⎪ n2 = xn + an x1 , · · · , xn−1 , x
x n2 ,
⎪
⎪ 2
⎪
⎪
⎪
⎪ 1
τ 1 1
⎪
⎪
x 2
= x + a x1 , · · · , x ,
x 2
, · · · ,
x 2
⎪
⎪
i i
2
i i i+1 n
⎪
⎪
⎪
⎪ - x i12
⎪
⎪ ∂ al 1
i−1
⎪
⎪ τ 1
⎪
⎪ − x1 , · · · , xi−1 , t, x i+1
2
,···,x
n2 d t, i = 2, · · · , n − 1,
⎪
⎪ 2 xi ∂ xl
⎪
⎪ l=1
⎪
⎪
⎪
⎪ 1
τ 1 1
⎪
⎨ x
12 = x1 + a1 x1 , x 22 , · · · , x
n2 ,
2
⎪
⎪ 1
τ 1 1
⎪
⎪ 1 = x
x 12 + a1 x 1 , x
22 , · · · , xn2 ,
⎪
⎪ 2
⎪
⎪
⎪
⎪ τ 1
⎪
⎪ j = x
1
j2 + aj x 1 , · · · , x
j , x
1
j+1
2
,···,x n2
⎪
⎪
x
⎪
⎪ 2
⎪
⎪
⎪
⎪ - j
j−1
∂ al 1
x
⎪
⎪ τ 1
⎪
⎪ + x1 , · · · , x
j−1 , t, x
j+1
2
,···,x
n2 d t, j = 2, · · · , n − 1,
⎪
⎪
1
2 x 2 ∂ xl
⎪
⎪ j l=1
⎪
⎪
⎪
⎪ 1
τ 1
⎩ x n2 + an x
n = x 1 , · · · , x
n−1 , x n2 .
2
(7.4)
Either (7.2) or (7.3) contains n − 1 implicit equations generally. But for fields a with
some specific properties, it will turn into explicit. For example,
∂ ai
= 0, i = 1, · · · , n (7.5)
∂ xi
It is easy to verify that if a = (a1 , · · · , an )T satisfies the condition (7.5), then the
scheme (7.6) is just the result of composing the Euler explicit schemes of the systems
associated to the fields a{k} (k = 1, · · · , n), i.e., we have
1G
τ
= Eaτ{n} ◦ · · · ◦ Eaτ{2} ◦ Eaτ{1} , (7.8)
where
Eaτ{k} = I + τ a{k} , k = 1, 2, · · · , n, I = identity. (7.9)
460 10. Volume-Preserving Methods for Source-Free Systems
{k}
In fact, Eaτ{k} are the phase flows eτa{k} , since a∗ a{k} = 0 for k = 1, 2, · · · , n,
which is implied by the condition (7.5). According to Theorem 6.3, we then get a 2nd
order explicit revertible volume-preserving scheme, with step transition map
τ τ τ τ τ τ
2G
τ 2{1} ◦ E
= Ea2{n} ◦ · · · ◦ Ea2{2} ◦ Ea2{1} ◦ E 2{2} ◦ · · · ◦ E
2{n}
a a a
= 1G 2
τ
τ2 = 2 gτ .
◦ 1G (7.10)
where Ess denotes the n × n matrix, of which only the entry at the s-th row and the
s-th column is equal to 1, and all other entries are 0. In this case, (8.2) and (8.3) have
much more simple forms. For example, for α = α(1,1) , (8.2) turns into
∂f1 ∂(f2 , · · · , fn )
= = 0, (8.5)
∂w1 ∂(w2 , · · · , wn )
Remark 8.4. For such a matrix α(1,1) , generating mapping f (w) of type α(1,1) , there
are n component f (w) = (f1 (w), f2 (w), · · · , fn (w)), in which n − 1 component
f2 (w), · · · , fn (w) is linear independent, satisfying condition
∂(f2 , · · · , fn )
∂(w2 , · · · , wn ) = 0,
We call (8.9) a Hamilton–Jacobi equation. The proofs of Theorems 8.1 and 8.2
can found in [Sha94b] .
Remark 8.6. If α = α(1,1) , then relations (8.9) and (8.10) turn into
∂ f1 ∂f
= −a1 (w1 , f2 , · · · , fn ) 1 , (8.11)
∂t ∂ w1
∂ fk ∂f
= ak (w1 , f2 , · · · , fn ) − a1 (w1 , f2 , · · · , fn ) k , k = 2, · · · , n, (8.12)
∂t ∂ w1
fk (w1 , · · · , wn , 0) = wk , k = 1, 2, · · · , n. (8.13)
n
∂ak
div a(z) = (z) = 0, z ∈ Rn , (8.14)
∂zk
k=1
for k ≥ 1, we have
10.8 Construction of Volume-Preserving Scheme via Generating Function 463
ip ≥1
1
k m
(k+1) 1 ∂ f (k) (w) 1
f (w) =− Cα a(E0 w) −
k+1 ∂w k+1 j!
m=1 j=1 i1 +···+ij =m
(k−m)
∂f (w)
· j
Cα Dα,E 0w
(Aα f (i1 ) (ω), · · · , Aα f (ij ) (w))
∂w
1
k
1
+ m
Aα Dα,E 0w
(Aα f (i1 ) (ω), · · · , Aα f (im ) (w)),
k+1 m!
m=1 i1 +···+im =k
ip 1
(8.20)
(k) (k)
where for ξ (k) = (ξ1 , · · · , ξn )T ∈ Rn (k = 1, 2, · · · , m), we get
⎡ ⎤
n
∂ m a1 (w)
⎢ ξα1 1 · · · ξαmm ⎥
⎢ ∂z α1
· · · ∂z αm ⎥
⎢ α1 ,···,αm =1 ⎥
⎢ .. ⎥
m
Dα,w (ξ (1) , · · · , ξ (m) ) = ⎢ . ⎥. (8.21)
⎢ ⎥
⎢ n
∂ m an (w) ⎥
⎣ ξ 1 · · · ξαmm ⎦
α ,···,α =1
∂zα1 · · · ∂zαm α1
1 m
Proof. Under the above proposition, if generating function f (w, t) = fα,a (w, t) is
dependent analytically on w and t in some neighborhood Rn for sufficient small t,
then it can be expressed as a power series
∞
f (w, t) = f (k) (w)tk .
k=0
By (8.10),
f (0) (w) = f (w, 0) = N0 w.
This is (8.19). Denote E0 = Aα N0 + B α = (Cα + Dα )−1 , then
∞
Aα f (w, t) + B α w = E0 w + Aα f (k) (w)tk .
k=1
* +
α α
∞
α (k) k
α(A f (w, t) + B w) = a E0 w + A f (w)t
k=1
∞
k 1 m
= a(E0 w) + tk Da,E0 w (Aα f (i1 ) (w), · · · , Aα f (im ) (w)).
m=1 i1 +···+im =k
m!
k=1
ip 1
(8.24)
for k ≥ 1,
(k) k−1
m
n
(k+1) 1 ∂f (w) 1 1
fi (w) = $1 (w) i∂w
k+1 a + k+1
1
m=1 j=1 i1 +···+ij =m α1 ,···,αj =2
j!
ip 1
(k−m)
∂ fi (w) ∂ a"1 (w)
j
(i )
· f (i1 ) (w) · · · fαjj (w)
∂ w1 ∂ wα1 · · · ∂ wαj α1
k n
1 1 ∂ m a"i (w) (i ) (im )
+ f 1 (w) · · · fαm (w),
k + 1 m=1 i m! ∂wα1 · · · ∂wαm α1
1 +···+im =k α1 ,···,αm =2
ip 1
i = 1, 2, · · · , n.
(8.27)
10.8.2 Construction of Volume-Preserving Schemes
In this subsection, we consider the construction of volume-preserving schemes[Sha94a]
for the source-free system (8.8). By Remark 8.3 of Theorem 8.1, for given time-
dependent scalar functions φ2 (w, t), · · · , φn (w, t) : Rn × R → R and C(w, " t) :
Rn−1 × R → R, we can get a time-dependent volume-preserving mapping g"(z, t). If
φ2 (w, t), · · · , φn (w, t) approximates the generating functions f2 (w, t), · · · , fn (w, t)
of the type α(1,1) of the source-free system (8.8), then suitable choice C(w, " t), g"(w, t)
approximates the phase flow gαt (z) = g(z, t). Fixing t as a time step, we can get a
difference scheme (volume-preserving schemes) whose transition from one time step
to the next is volume-preserving. By Remark 8.8 of Theorem 8.7, generating functions
f2 (w, t), · · · , fn (w, t) can be expressed as power series. So, a natural way to approx-
imate f2 (w, t), · · · , fn (w, t) is take the truncation of the series. However, we have to
choose a suitable C(w, " t) in (8.7) to guarantee the accuracy of the scheme.
Assume that
m
(k)
φmi (w, t) = fi (w)tk , i = 2, · · · , n (8.28)
k=0
10.8 Construction of Volume-Preserving Scheme via Generating Function 465
and
m
(k)
ψ1m (w, t) = f1 (w)tk . (8.29)
k=0
and
-
w1 ∂ (φ(m) , · · · , φn )
(m)
(m)
φ1 (w, t) = C (m) (w2 , · · · , wn , t) +
∂ (w2 , · · · , wn ) (ξ, w2 , · · · , wn , t)d ξ,
2
w1,0
(8.31)
then we have,
Theorem 8.9. Using Theorem 8.5 and Theorem 8.7 for sufficiently small τ ≥ 0
(m) (m)
as the time step, defining mapping φ(m) (w, τ ) = (φ1 (w, τ ), φ2 (w, τ ), · · · ,
(m) (m)
φn (w, τ ))T with the components φi (w, τ )(i = 1, 2, · · · , n) given as above for
m = 1, 2, · · · , then the mapping
= φ(m) (w, τ ),
w −→ w (8.32)
i = 2, · · · , n, (8.33)
⎩ (m)
zik+1 = φi (z1k+1 , z2k , · · · , znk , τ ),
By Theorem 8.1, Remark 8.3, Remark 8.4, and Equation (8.31), the relation (8.33)
defines a time-dependent volume-preserving z = z k → z k+1 = z = g"(z, τ ). That is,
(8.33) is a volume-preserving scheme.
Noting that
(m)
φi (w, τ ) = fi (w, τ ) + O(τ m+1 ), i = 2, · · · , n,
(m)
ψ1 (w, τ ) = f1 (w1 , τ ) + O(τ m+1 ),
466 10. Volume-Preserving Methods for Source-Free Systems
Remark 8.11. We can get volume-preserving scheme similar to the above one if we
consider the types α = α(s,s) (2 ≤ s ≤ n), instead of α = α(1,1) .
(1) k+1 k i = 2, · · · , n,
⎪
⎩ zi = φi (z1 , z2 , · · · , zn , τ ),
k k
where
(1)
φ1 (w, τ ) = −τ a1 (0, w2 , · · · , wn )
∂ a2 ∂ a2 ∂ a2
1+τ τ ··· τ
∂ w 2 ∂ w3 ∂ wn
∂ a3 ∂ a3 ∂ a3
- w1 τ 1+τ ··· τ
∂ w 2 ∂ w3 ∂ wn
+ (ξ, w2 , · · · , wn )d ξ,
0 .. .. ..
. . .
∂ an ∂ an ∂ an
τ τ ··· 1+τ
∂ w2 ∂ w3 ∂ wn
(1)
φi (w, τ ) = wi + τ ai (w).
⎪ (2)
k+1
⎩ zi = φi (z1k+1 , z2k , · · · , znk , τ ), i = 2, · · · , n,
10.9 Some Volume-Preserving Algorithms 467
where
-
w1 ∂ (ψ2(2) , · · · , ψn(2) )
(2) (2)
ψ1 (0, w2 , · · · , wn , τ )
φ1 (w, τ ) = + ∂ (w2 , · · · , wn ) (ξ, w2 , · · · , wn )d ξ,
0
(2) (2)
φi (w, τ ) = ψi (w, τ ), i = 2, · · · , n,
and
(2) (2) T 1 ∂" a(w)
ψ (2) (w, τ ) = ψ1 (w, τ ), · · · , ψn (w, τ ) = w + τ"
a(w) + τ 2 "
a(w),
2 ∂ w1
T
"
a(w) = − a1 (w), a2 (w), · · · , an (w) .
In this section, we analyze and study under conditions a source-free system that has
volume-preserving R–K schemes.
Obviously, this is a source-free system. Its phase flow in Rp+q preserves the phase
volume of (p + q) form
d x1 ∧ d x2 ∧ · · · ∧ d xp ∧ d y1 ∧ d y2 ∧ · · · ∧ d yq .
Only R–K and P–R–K are to be discussed. We wish, some of the phase volume is
preserved.
The formula of a general m-th stage P–R–K method with time step h applied to
system (9.1) is read as
468 10. Volume-Preserving Methods for Source-Free Systems
m
m
ξi = xn + h dij g(ηj ), ηi = yn + h cij f (ξj ), 1 ≤ i ≤ m,
j=1 j=1
m
xn+1 = xn + h δj g(ηj ),
(9.2)
j=1
m
yn+1 = yn + h γj f (ξj ),
j=1
For arbitrary 1 ≤ r ≤ m and two arbitrary ordered sets (k1 , · · · , kr ) and (l1 , · · · , lr )
of different natural numbers from (1, m), dij and cij are elements (i, j) with respect
to matrix D− and C − .
Next, for system (9.1), we construct some volume-preserving method by P–R–K
method, using the above criteria.
First we consider volume-preserving by R–K method for linear system.
Linear system of ODE is read as
ẏ = M y, (9.3)
different classes and find out whether there are volume-preserving R–K method in any
class.
Now, we need the following notations:
A = A ⊗ En , M = diag (M, M, · · · , M ) = Es ⊗ M,
b = bT ⊗ En , Y = [Y1 , Y2 , · · · , Ys ]T (9.5)
y n = [yn , yn , · · · , yn ]T , e = e ⊗ es ,
So,
yn+1 = (En + (hM b(I − hM A)−1 e)yn
∂yn+1
=⇒ = En + hM b(I − hM A)−1 e. (9.7)
∂yn
Lemma 9.2. Let A, D be non-degenerate m × m and n × n matrices respectively and
B an m × n and C an n × m matrix, then
det A det (D + CA−1 B) = det D det (A + BD−1 C). (9.8)
The proof can be found in any textbook of linear algebra.
By Lemma 9.2, it is easy to get from (9.7)
% &
∂ yn+1 det (I − hM A − eM b)
det = .
∂ yn det (I − hM A)
Now, if (9.10) is identical to 1, we arrive at the criterion for R–K method (9.4) to be
volume-preserving scheme as
det (λI − N − ) = det (λI − N ), ∀ λ ∈ R. (9.11)
470 10. Volume-Preserving Methods for Source-Free Systems
Theorem 9.3. If dimension of M is odd, then all the R–K methods based on high or-
der quadrature formula such as Gauss–Legendre, Radau, and Lobatto are not volume-
preserving.
Proof. Note that N = A ⊗ M and N − = A− ⊗ M . If the method is volume-
preserving, then
det N = det(N − ) ⇐⇒ det(A ⊗ M ) = det(A− ⊗ M )
⇐⇒ (det A)n (det(M ))s = (det(A− ))n (det M )s
⇐⇒ (det A)n = (det(A− ))n
⇐⇒ det A = det(A− ). (9.12)
Now, we need the W -transformation proposed by Hairer and Wanner[HW81] . They
introduced a generalized square matrix W defined by
W = (p0 (c), p1 (c), · · · , pn−1 (c)), (9.13)
where the normalized shifted Legendre polynomials are defined by
√
k k k + i
pk (x) = 2k + 1 (−1)k+i xi , k = 0, 1, · · · , s − 1. (9.14)
i i
i=0
1
where ξk = √ (k = 0, 1, · · · , s − 1).
2 4k2 − 1
However, X − = W −1 A− W , then
⎡ 1 ⎤
− −ξ1
⎢ 2 ⎥
⎢ ξ 0 −ξ2 ⎥
⎢ 1 ⎥
− ⎢ ξ ⎥
X =⎢ 2 ⎥.
⎢ .. .. ⎥
⎢ . . −ξs−1 ⎥
⎣ ⎦
ξs−1 0
Theorem 9.4. [QL00] If the dimension of M is even, then the R–K methods based on
high order quadrature formulas such as Gauss–Legendre, Lobatto III A, Lobatto III
B, Lobatto III S, Radau IB, and Radau IIB are volume-preserving, iff
So, the R–K method based on high order quadrature formula (Gauss–Legendre,
Lobatto IIIA, Lobatto IIIB, Lobatto IIIS, Radau IB, and Radau IIB) are volume-
preserving. The Theorem 9.4 says that for the methods to preserve volume, the system,
in some sense, must be similar to a Hamiltonian system. If the matrix M similar to
an infinitesimally symplectic matrix, i.e., there is an invertible matrix P , subjected
to P −1 M P = JS, S T = S, then we can transform the system to a Hamiltonian
system by a coordinate transformation. In this situation, the volume-preserving R–K
methods and the symplectic R–K methods almost have no difference, that is, if P is a
symplectic matrix, then volume-preserving R–K methods are equivalent to symplec-
tic R–K methods; and in this case, they can be transformed to one another by a linear
transformation.
In the case r = 1, if the necessary and sufficient condition of Lemma 9.5 are sat-
isfied, then a 2-stage P–R–K method is volume-preserving. This condition is the
same condition of symplecity on the class of separable Hamiltonian system. Thus
for system (9.3), all 2-stage P–R–K methods proposed in[Sun95] are volume-preserving
algorithms[QL00] .
1 1
0 − 0 0 0
4 4
2 1 5 2 1 1
3 4 12 3 3 3
1 3 1 3
4 4 4 4
1 5 1 1 1
− 0
3 12 12 3 3
3 1
1 1 1 0
4 4
3 1 3 1
4 4 4 4
1 1
0 − 0 0 0
2 2
1 1
1 1 1 0
2 2
1 1 1 1
2 2 2 2
√ √
1 3 1 + 2σ 1 − 2σ 3
− −
2 6 4 4 6
√ √
1 3 1 − 2σ 3 1 + 2σ
+ +
2 6 4 6 4
1 1
2 2
√ √
1 3 1 − 2σ 1 + 2σ 3
− −
2 6 4 4 6
√ √
1 3 1 + 2σ 3 1 − 2σ
+ +
2 6 4 6 4
1 1
2 2
10.9 Some Volume-Preserving Algorithms 473
m
m
m
ξi = xn + h dij g(ηj ), ηi = yn + h cij h(wj ), wj = zn + h eij f (ξj ),
j=1 j=1 j=1
(9.17)
m
m
m
xn+1 = xn +h αj g(ηj ), yn+1 = yn +h βj h(wj ), zn+1 = zn +h γj (ξj ).
j=1 j=1 j=1
1 1
0 − 0 0 0 a a 0
2 2
1 1
1 1 1 0 b+c b c
2 2
1 1 1 1
b 2c
2 2 2 2
Suitably choose a, b, c, as method can get global truncation error with order O(h2 ).
Remark 9.13. Theorem 9.10 can be extended with no difficulty to the following sys-
tem:
ẋ1 = f2 (x2 ) ẋ2 = f3 (x4 ), ···, ẋn = f1 (x1 ). (9.19)
474 10. Volume-Preserving Methods for Source-Free Systems
γi dij + δi cji − γi δj = 0,
we can say this integrator is symplectic. If system is not Hamiltonian, we cannot say
that this P–R–K method is symplectic. The main problem is that we say a scheme
is symplectic because it preserves symplectic structure for a given system. Therefore,
only Hamiltonian system possesses symplectic structure. Consequently, we cannot say
“volume-preserving P–R–K methods form a subset of symplectic ones”.
Until now, we gave some criteria for volume-preserving by R–K and P–R–K
methods. In fact, it is almost impossible based on these criteria to construct volume-
preserving algorithm with high order accuracy. Indeed, we even cannot predict that
there exists schemes which satisfied those criteria. We are too far to resolve these
problems.
It should be noted that in the above discussion, we always suppose system is not
reducible. In other words, det M = 0. But in practice, some systems are reducible, for
example
ẋ = cy − bz,
ẏ = az − cx, a, b, c ∈ R.
ż = bx − ay,
Theorem 9.14. If the dimension of M is odd, then the R–K methods based on high
order quadrature formulae, such as LobattoIIIA, LobattoIIIB, LobattoIIIS, RadauI,
RadauIIB etc., are volume-preserving, iff
λ(M ) = λ1 , λ2 , · · · , λ n2 , 0, −λ1 , −λ2 , − · · · , −λ n2 .
∂ fi
schemes is volume-preserving iff the Jacobian = M is, in some sense, similar to
∂ yi
an infinitesimally symplectic matrix. That is, the eigenvalues of M can be specified as
λ(M ) = λ1 , λ2 , · · · , λ n2 , −λ1 , −λ2 , − · · · , −λ n2 ,
or
λ(M ) = λ1 , λ2 , · · · , λ n2 , 0, −λ1 , −λ2 , − · · · , −λ n2 .
Bibliography
[DV84] K. Dekker and J.G. Verwer: Stability of Runge–Kutta Methods for Stiff Initial Value
Problems. Elesevier Science Pub. B. V., North-Holland, Amsterdam, (1984).
[FS95] K. Feng and Z. J. Shang: Volume-preserving algorithms for source-free dynamical
systems. Numer. Math., 71:451–463, (1995).
[FW94] K. Feng and D.L. Wang: Dynamical systems and geometric construction of algo-
rithms. In Z. C. Shi and C. C. Yang, editors, Computational Mathematics in China, Con-
temporary Mathematics of AMS, Vol. 163, pages 1–32. AMS, (1994).
[HW81] E. Hairer and G. Wanner: Algebraically stable and implementable Runge–Kutta meth-
ods of high order. SIAM J. Numer. Anal., 18:1098–1108, (1981).
[MQ04] R.I. McLachlan and G.R.W. Quispel: Explicit geometric integration of polynomial
vector fields. BIT, 44:513–538, (2004).
[QD97] G. R. W. Quispel and C. P. Dyt: Solving ODE’s numerically while preserving sym-
metries, Hamiltonian structure, phase space volume, or first integrals. In A. Sydow, editor,
Proceedings of the 15th IMACS World Congress, pages 601–607. Wissenschaft & Technik,
Berlin, (1997).
[QD98] G. R. W. Quispel and C. P. Dyt: Volume-preserving integrators have linear error
growth. Physics Letters A, 202:25–30, (1998).
[QL00] M. Z. Qin and H. W. Li: Volume preserving R–K methods for linear systems. Acta
Applicandae Mathematicae, 16:430–434, (2000).
[QM03] G. R. W. Quispel and D. I. McLaren: Explicit volume-preserving and symplectic
integrators for trigonometric polynomial flows. J. of Comp. Phys., 186(1):308–316, (2003).
[Qui95] G. R. W. Quispel: Volume-preserving integrators. Physics Letters A, 206:26–30,
(1995).
[QZ92] M. Z. Qin and W. J. Zhu: Construction of higher order symplectic schemes by com-
position. Computing, 47:309–321, (1992).
[QZ93] M. Z. Qin and W. J. Zhu: Volume-preserving schemes and numerical experiments.
Computers Math. Applic., 26:33–42, (1993).
[Sco91] C. Scovel: Symplectic numerical integration of Hamiltonian systems. In T. Ratiu,
editor, The Geometry of Hamiltonian Systems, pages 463–496. Springer, New York, (1991).
[Sha94a] Z Shang: Construction of volume-preserving difference schemes for source-free sys-
tems via generating functions. J. Comput. Math., 12:265–272, (1994).
[Sha94b] Z. Shang: Generating functions for volume-preserving mappings and Hamilton–
Jacobi equations for source-free dynamical systems. Science in China (series A), 37:1172–
1188, (1994).
[Sun95] G. Sun: Construction of high order symplectic Partitioned–Runge–Kutta methods. J.
Comput. Math., 13(1):40–50, (1995).
[Sur96] Y. B. Suris: Partitioned–Runge–Kutta methods a phase volume preserving integrators.
Physics Letters A, 220:63–69, (1996).
[TH85] A. Thyagaraja and F.A. Haas: Representation of volume-preserving maps induced by
solenoidal vector fields. Phys. Fluids, 28:1005, (1985).
[Wey40] H. Weyl: The method of orthogonal projection in potential theory. Duke Math. J.,
7:411–444, (1940).
Chapter 11.
Contact Algorithms for Contact Dynamical
Systems
z z
6 6
........ .... ..
... ............................... ... ..
... ................ ... ..
...
.. ................
................ ... ..
...
V
.. N ...... ... .....
.
.... ... ..
.. ...
. ..
...... .. ... .. ..
..
.. .. ... ..
.. ..
. ..
... .. ... .. ..
.... . ... .. ..
..
.... .. ......
. ..
..
.. ..
...... ... .
........ ... .. ..
..
. . ...... . ..
.... .. ..... ... .. ...
.... ... ...
..
.. N ...
.... .. .. ...
....... ...
. ..
..
...
....
.
........................ ... .. ..
..........
........ .. ...
.. ..
........
.......
...... ... .... V ...
.
-y - y
O ...... ... O
......
...... ...
...... ..
M ..... ...
...... ..
........
.
M
x / x /
Fig. 1.1. Meaning of definition
..
...
...
...
...
.
........
z ....
.....
. .
V
........
..................... ......................
...
...
...
... ......................... ...
.............
6
...
... .............. ..
... ...................... ...
...... .... .. . . ... .. ......
... ..................................... ...
... ............
............. ..
.. ..
... ............. ..
... .............. k+1 ..
.. .............
.......
.........
V ..
..
..
..
.. k
..
. Γ
...
..
..
- y
M
x /
Fig. 1.2. Integral surface with initial manifold of Γ
......
... ...............................
... ................
................
z 6 ... ................
.. .. ........
.. . ...
....... 6 ...
...
.. ..
. ...
.... ...
.... ...
.... ...
...... ...
.... ...
.... ...
... ..
....... ...
....................... .
.......... ...
........
........ ...
.......
.......
......
...... ...
...
...
- y
...... ...
......
..... ...
...... ....
.......
x /
Fig. 1.3. Hyperplane
and [0, x, 1] not all equal to zero, it is defined as a 2-dimensional field of hyperplane.
When x = 0,
⎡ ⎤T ⎡ ⎤
0 ηx
⎢ ⎥ ⎢ ⎥
⎣ 0 ⎦ ⎣ ηy ⎦ = 0.
1 0
Each point with a hyperplane intersecting wall defines a direction field, see Fig. 1.4
and 1.5.
Next, we prove that in R3 space, there does not exist an integral surface which
can be given by the 1-form α = xd y + d z, where x, y is horizontal coordinate, z is
vertical coordinate, see Fig. 1.6.
Consider a pair of vectors emanating from the origin (0,0,0) and lying in the hori-
zontal plane of our coordinate systems; another integral curve from (0,0,0) to (0,1,0),
and then from (0,1,0) to (1,1,0), and another integral curve from (0,0,0) to (1,0,0),
and then from (1,0,0) to (1,1, −1). As a result, these two curves cannot close up. The
difference in the heights of these points is 1, this difference can be considered as a
measure of the nonintegrability of the field.
We have four direction fields from the origin point 0 to walls of east, south, west,
and north, respectively, describing by Fig. 1.5.
480 11. Contact Algorithms for Contact Dynamical Systems
- : :
-
~
- : : ~
-
: : ~
~
Definition 1.3. A differential 1-form α which is nowhere equal to the zero form on
a manifold M is called a contact form if the exterior derivative dα of α defines a
nondegenerate exterior 2-form in every plane α = 0.
11.1 Contact Structure 481
Example 1.4. Consider the space R2n+1 with the contact structure by the 1-form
α = d u + p d q. Where q = (q1 , · · · , qn ), u, p = (p1 , · · · , pn ), α is not equal to zero
form at any point in R2n+1 , and consequently defines the field of 2n-dimensional
planes α = 0 in R2n+1 .
Example 1.5. The form constructed in Example 1.4 is a contact form, the exterior
derivatives of the form α is equal to
d α|α=0 = d q1 ∧ d p1 + · · · + d qn ∧ d pn .
here we have used 3-symbol notation to denote the coordinates and vectors on R2n+1
d
ẋ = a(x, y, z), ẏ = b(x, y, z), ż = c(x, y, z), · =: , (1.3)
dt
where the contactivity condition of the vector field f is
Lf α = λf α, (1.4)
Lf α = if d α + d if α. (1.5)
It is easy to show from (1.4) and (1.5) that to any contact vector field f on R2n+1 ,
there corresponds a function K(x, y, z), called contact Hamiltonian, such that
11.1 Contact Structure 483
a = −Ky + Kz x, b = Kx , c = K − xT Kx =: Ke . (1.6)
In fact, (1.6) represents the general form of a contact vector field. Its multiplier, de-
noted as λf from now, is equal to Kz .
z z(x, y, z)
t : R
2n+1
for some everywhere non-vanishing function μgK → R. Moreover, we have
the following relation between μG∗k and the Hamiltonian K:
- t
μgK
t = exp (Kz ◦ gK
s
)d s. (1.11)
0
For general contact systems, condition (1.10) is stringent for algorithmic approx-
imations to phase flows because only the phase flows themselves satisfy it. We will
construct algorithms for contact systems such that the corresponding algorithmic ap-
proximations to the phase flows satisfy the condition (1.10), of course, probably, with
different, but everywhere non-vanishing, multipliers from μgK t . We call such algo-
Consider ( )
R2n+2
+ = (p0 , p1 , q0 , q1 ) ∈ R2n+2 | p0 > 0 (2.2)
as a conic symplectic space with the standard symplectic form
F ◦ Tλ = Tλ ◦ F, ∀ λ > 0, (2.5)
P0 d Q0 + P1 d Q1 = p0 d q0 + p1 d q1 , or P d Q = pd q, (2.9)
QT
p · P = 0, QT
q · P = p. (2.10)
if and only if
xi fxi (x1 , x2 , · · · , xn ) = kf (x1 , x2 , · · · , xn ).
QT T
p Pp − Pp Qp = O, QT T
q Pq − Pq Qq = O, QT T
q Pp − Pq Qp = I. (2.12)
p = QT T T T T T
q Pp p − Pq Qp p = Qq P, O = Qp Pp p − Pp Qp p = Qp P. (2.13)
P = Pp QT T
q P − Pq Qp P = Pp p,
(2.15)
0 = Qq QT T
p P − Qp Qq P = Qq p.
b. Consider R+ × R2n+1 as the product of the positive real space R+ and the
contact space R2n+1 . We use (w, x, y, z) to denote the coordinates of R+ × R2n+1
with w > 0 and with x, y, z as before.
Definition 2.4. A map G: R+ × R2n+1 → R+ × R2n+1 is called a positive product
map if it is composed by a map g : R2n+1 → R2n+1 and a positive function γ :
R2n+1 → R+ in the form
⎡ ⎤ ⎡ ⎤
w W
⎢ ⎥ ⎢ ⎥
⎢ x ⎥ ⎢ X ⎥
⎢ ⎥ −→ ⎢ ⎥, W = w γ(x, y, z), (X, Y, Z) = g(x, y, z). (2.16)
⎣ y ⎦ ⎣ Y ⎦
z Z
P1 (1, x, z, y)
X= , Y = Q1 (1, x, z, y), Z = Q0 (1, x, z, y). (2.20)
P0 (1, x, z, y)
and compute
((0, 0, P0T , P1T )F∗ − (0, 0, pT T
0 , p1 )) ◦ S S∗
= (0, 0, P0T , P1T ) ◦ S (F∗ ◦ S)S∗ − (0, 0, pT T
0 , p1 ) ◦ S S∗
= (0, 0, wγ, wγX T )(F∗ ◦ S)S∗ − (0, 0, w, wxT )S∗
= (0, 0, wγ, wγX T )(S∗ ◦ G)G∗ − (0, 0, w, wxT )S∗
= wγ 0, (0, X T , 1)g∗ − wγ 0, γ −1 (0, xT , 1) .
Lemma 2.5 establishes correspondences between conic symplectic space and con-
tact space and between conic symplectic maps and contact maps. We call the transform
from F to G = S −1 ◦ F ◦ S = γ ⊗ g contactization of conic symplectic maps, the
transform from G = γ ⊗ g to F = S ◦ GS −1 symplectization of contact maps and
call the transform S : R+ × R2n+1 → R2n+1+ symplectization of contact space, and
the transform C = S −1 : R2n+2
+ → R + × R 2n+1
contactization of conic symplectic
space.
488 11. Contact Algorithms for Contact Dynamical Systems
1
C= (I + JB), B = B T ∈ Sm(2n + 2), (3.1)
2
establishes a 1-1 correspondence between near-zero Hamiltonian vector fields z →
a(z) ≡ J∇φ(z) and near-identity symplectic maps z → g(z) via generating relation
and combining Lemma 3.1 and Lemma 3.2, we find that matrix
5 6
C0 O
C= , C0 ∈ gl(n + 1), (3.3)
O I − C0T
5 6 5 6
p0 q0
where p = and q = are given by
p1 q1
⎧
⎪
⎪ p0 + (1 − α)p0 + β T (
p0 = α p1 − p1 ),
⎪
⎪
⎪
⎪
⎨ p1 = δ
p1 + (I − δ)p1 + γ(
p0 − p0 ),
(3.5)
⎪
⎪ q 0 = (1 − α)q0 + αq0 − γ T (q1 − q1 ),
⎪
⎪
⎪
⎪
⎩
q 1 = (I − δ T )
q1 + δ T q1 − β(
q0 − q0 ).
i.e.,
p1
where x = , y = q1 , z = q0 on the right hand side. So, under contactizing
p0
transforms
⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤ ⎡ ⎤
w p0 w
w p0
w
⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ x ⎥ ⎢ p ⎥ ⎢ wx ⎥ ⎢ x ⎥ ⎢ p ⎥ ⎢ w ⎥
⎢ ⎥ ⎢ 1 ⎥ ⎢ ⎥ ⎢ ⎥ ⎢ 1 ⎥ ⎢ x ⎥
S: ⎢ ⎢
⎥ −→ ⎢
⎥ ⎢
⎥=⎢
⎥ ⎢
⎥,
⎥
⎢
⎢
⎥ −→ ⎢
⎥ ⎢
⎥=⎢
⎥ ⎢
⎥,
⎥
⎢ y ⎥ ⎢ q0 ⎥ ⎢ z ⎥ ⎢ y ⎥ ⎢ q0 ⎥ ⎢ z ⎥
⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦ ⎣ ⎦
z q1 y z q1 y
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
w p0 w
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ x ⎥ ⎢ ⎥ ⎢ wx ⎥
⎢ ⎥ ⎢ p1 ⎥ ⎢ ⎥
⎢ ⎥ −→ ⎢ ⎥=⎢ ⎥, (3.8)
⎢ ⎥ ⎢ ⎥ ⎢ ⎥
⎢ y ⎥ ⎢ q0 ⎥ ⎢ z ⎥
⎣ ⎦ ⎣ ⎦ ⎣ ⎦
z q1 y
⎧
⎪
⎪ − w = −wψz (x, y, z),
w
⎪
⎪
⎪
⎪
⎨ w
x − wx = −wψy (x, y, z),
(3.9)
⎪
⎪ z − z = ψe (x, y, z),
⎪
⎪
⎪
⎪
⎩
y − y = ψx (x, y, z),
Since the p0 -axis is distinguished for the contactization in which we should always
w p w p
= = 0 and μ = = 0 ,
take p0 = 0, it is natural to require β = 0 in (3.5). Let μ
w p0 w p0
we obtain from Equations (3.9) and (3.10)
1 + αψz (x, y, z)
=
μ , μ = 1 + αψz (x, y, z), (3.11)
1 − (1 − α)ψz (x, y, z)
and the induced contact transformation on the contact (x, y, z) space R2n+1 is
⎧
⎪ x − x = −ψy (x, y, z) + ψz (x, y, z) (1 − α)x + αx ,
⎪
⎨
y − y = ψx (x, y, z), (3.12)
⎪
⎪
⎩
z − z = ψe (x, y, z),
where ⎧
⎪ d1 = I − (1 − α)ψz (x, y, z) δ,
⎪
⎪
⎨
d2 = I + αψz (x, y, z) (I − δ), (3.14)
⎪
⎪
⎪
⎩
d0 = −ψz (x, y, z)γ.
Summarizing the above discussions, we have:
Theorem 3.3. Relations (3.12) – (3.14) give a contact map (x, y, z) → (x, y, z) via
α O
contact generating function ψ(x, y, z) under the type C0 = . Vice versa.
γ δ
11.3 Contact Generating Functions for Contact Maps 491
However, the difficulty in the algorithmic implementation lies in the fact that, un-
like ȳ and z, which are linear combinations of y, y and z, z with constant matrix coef-
ficients, since x = d1 x + d2 x + d0 and d1 , d2 are matrices with coefficients depending
on ψ̄z = ψz (x, y, z) which in turn depends on (x, y, z) the combination of x from x
and x is not explicitly given, the entire equations for solving x, y, z in terms of x, y, z
are highly implicit. The exceptional cases are the following:
(E1) α = 0, δ = On , γ = O,
= 1 − ψz (x, y, z), μ = 1,
μ (3.15)
⎧
⎪ x − x = −ψ y (x, y , z) + x
ψz (x, y, z),
⎪
⎨
y − y = ψx (x, y, z), (3.16)
⎪
⎪
⎩
z − z = ψe (x, y, z) = ψ(x, y, z) − xT ψx (x, y, z).
(E2) α = 1, δ = In , γ = O,
= μ = 1 + ψz (
μ x, y, z), (3.17)
⎧
⎪ x − x = −ψy ( x, y, z) + xψz (
x, y, z),
⎪
⎨
y − y = ψx (
x, y, z), (3.18)
⎪
⎪
⎩
z − z = ψe (
x, y, z) = ψ(x, y, z) − xT ψx (
x, y, z).
1 1
(E3) α = , δ = In , γ = O,
2 2
1
1+ψz (x, y, z) 1
=
μ 2 , μ = 1 + ψz (x, y, z), (3.19)
1 2
1 − ψz (x, y, z)
2
⎧ +x
x
⎪
⎪ − x = −ψy (x, y, z) + ψz (x, y, z)
x ,
⎪
⎨ 2
y − y = ψx (x, y, z), (3.20)
⎪
⎪
⎪
⎩
z − z = ψe (x, y, z) = ψ(x, y, z) − xT ψx (x, y, z),
with
+x
x 1 y + y z + z
x= − ψz (x, y, z)(
x − x), y= , z= . (3.21)
2 4 2 2
with
+x
x 1 y + y z + z
x= − λ (
x − x), y= , z= . (3.24)
2 4 2 2
Note that the symplectic map induced by generating function φ from the relation
(3.2) can be represented as the composition of the maps, non-symplectic generally,
z → z and z → z
z = z + CJ∇φ(z),
z = z + (I − C)J∇φ(z).
Theorem 3.4. Contact map (x, y, z) → ( x, y, z) induced by contact generating func-
tion ψ from the relations (3.12)–(3.14) can be represented as the composition of the
maps (x, y, z) → (x, y, z) and (x, y, z) → ( x, y, z) which are not contact generally
and given, respectively, as follows
⎧
⎪ x − x = −δψy (x, y, z) + αψz (x, y, z)x − γψz (x, y, z),
⎪
⎨
y − y = (I − δ T )ψx (x, y, z), (3.25)
⎪
⎪
⎩
z − z = (1 − α)ψe (x, y, z) − γ T ψx (x, y, z)
and
⎧
⎪ x − x = −(I − δ)ψy (x, y, z) + (1 − α)ψz (x, y, z)
x + γψz (x, y, z),
⎪
⎨
y − y = δ T ψx (x, y, z), (3.26)
⎪
⎪
⎩
z − z = αψe (x, y, z) + γ T ψx (x, y, z).
(3.25) and (3.26) are the 2-stage form of the generating relation (3.12) of the contact
α O
map induced by generating function ψ under the type C0 = . Correspond-
γ δ
ing to the exceptional cases (E1), (E2) and (E3), the above 2-stage representation has
simpler forms, we no longer use them here.
Consider contact system (1.3) with the vector field a defined by contact Hamiltonian
K according to Equation (1.6). Take ψ(x, y, z) = sK(x, y, z) in (3.12) – (3.14) as the
generating function, we then obtain contact difference schemes with 1st order of ac-
α O
curacy of the contact system (1.3) associated with all possible types C0 = .
γ δ
The simplest and important cases are (write K x = Kx (x, y, z), etc.) as follows[Fen95] .
11.4 Contact Algorithms for Contact Systems 493
= x + sa(
x x, y, z), y = y + sb(
x, y, z), z = z + sc(
x, y, z).
, i.e., x
It differs from (4.2) only in one term for x K(x, y, z) instead of xK(x, y, z).
This minute, but delicate, difference makes (4.2) contact and the other non-contact!
It should be noted that the Q and P methods are of order one of accuracy and the
C method is of order two. The proof is similar to that for symplectic case. In principle,
one can construct the contact difference schemes of arbitrarily high order of accuracy
for contact systems, as was done for Hamiltonian systems, by suitably composing the
Q, P or C method and the respective reversible counterpart[QZ92] . Another general
method for the construction of contact difference schemes is based on the generat-
ing functions for phase flows of contact systems which will be developed in the next
section.
∂ t
φ (u) = H u + (I − C)J∇φt (u) , with u = (p0 , p1 , q0 , q1 )T , (5.1)
∂t
⎡ ⎤ ⎡ ⎤
p0 − (1 − α)φq0 p0 (1 − (1 − α)ψz )
⎢ ⎥ ⎢ ⎥
⎢ p1 + γφq − (I − δ)φq ⎥ ⎢ p0 (x + γψz − (I − δ)ψy ) ⎥
⎢ 0 1 ⎥ ⎢ ⎥
u + (I − C)J∇φt (u) = ⎢
⎢
⎥=⎢
⎥ ⎢
⎥
⎥
T
⎢ q0 + αφp0 + γ φp1 ⎥ ⎢ z + αψe + γ T ψx ⎥
⎣ ⎦ ⎣ ⎦
q1 + δ T φp1 y + δ T ψx
and
H(u + (I − C)J∇φt (u))
% &
x − (I − δ)ψy + γψz
= p0 1 − (1 − α)ψz K , y + δ T ψx , z + αψe + γ T ψx .
1 − (1 − α)ψz
% &
∂ t x + δψy + γψz
ψ = (1+αψz )K , y + (δ T − I)ψx , z + (α − 1)ψe + γ T ψx .
∂t 1 + αψz
(5.5)
(5.3) and (5.5) define the same function ψ t . When t = 0, etK = I, so we should
impose the initial condition
ψ 0 (x, y, z) = 0, (5.6)
for solving the first order partial differential equation (5.3) or (5.5). We call both equa-
tions the Hamilton–Jacobi equations of the contact! system associated with the contact
α O
Hamiltonian K and the matrix C0 = γ δ
.
Specifically, we have Hamilton–Jacobi equations for particular cases:
(E1) α = 0, δ = O, γ = O.
% &
∂ t x − ψyt
ψ = (1 − ψz )K
t
, y, z = K(x, y − ψxt , z − ψet ). (5.7)
∂t 1 − ψzt
(E2) α = 1, δ = In , γ = O.
1
= u + (I − C)J∇φt (u) and ū = u − CJ∇φt (u), then we have
Proof of the claim: let u
u = Cu + (I − C)ū.
496 11. Contact Algorithms for Contact Dynamical Systems
% &
∂ t x + ψyt
ψ = K(x, y + ψxt , z + ψet ) = (1 + ψzt )K , y, z . (5.8)
∂t 1 + ψzt
1 1
(E3) α = , δ = In , γ = O.
2 2
1 t
⎛ ⎞
ψy x−
∂ t 1 t 1 t 1
ψ = 1 − ψz K ⎝ 2 ,y + ψ , z + ψet ⎠
∂t 2 1 2 x 2
1 − ψzt
2
⎛ 1
⎞
x + ψyt
1 t ⎝ 1 t 1
= 1 + ψz K 2 ,y − ψ , z − ψet ⎠ . (5.9)
2 1 2 x 2
1 + ψzt
2
Remark 5.1. On the construction of high order contact difference schemes.
If K is analytic, then one can solve ψ t (x, y, z) from the above Hamilton–Jacobi
equations in the forms of power series in time t. Its coefficients are recursively de-
termined by the K and the related matrix C0 . The power series are simply given
from the corresponding conic Hamiltonian generating functions φt (p0 , p1 , q0 , q1 ) by
ψ t (x, y, z) = ψ t (1, x, z, y), since the power series expressions
of φt with respect to
p1
t from the conic Hamiltonian H(p0 , p1 , q0 , q1 ) = p0 K , q1 , q0 have been well
p0
given in[FW94] . Taking a finite truncation of the power series up to order m, an arbi-
trary integer, with respect to the time t and replacing by the truncation the generating
function ψ in (3.12)–(3.14), then one obtains a contact difference scheme of order
m for the contact system defined by the contact Hamiltonian K. The proofs of these
assertions are similar to those in the Hamiltonian system case, hence are omitted here.
Bibliography
[Arn78] V. I. Arnold: Ordinary Differential Equations. The MIT Press, New York, (1978).
[Arn88] V. I. Arnold: Geometrical Methods In The Theory of Ordinary Differential Equations.
Springer-Verlag, Berlin, (1988).
[Arn89] V. I. Arnold: Mathematical Methods of Classical Mechanics. Springer-Verlag, GTM
60, Berlin Heidelberg, Second edition, (1989).
[Etn03] J. Etnyre: Introductory lectures on contact geometry. In Proc. Sympos. Pure Math,
volume 71, page 81C107. SG/0111118, (2003).
[Fen93b] K. Feng: Symplectic, contact and volume preserving algorithms. In Z.C. Shi and
T. Ushijima, editors, Proc.1st China-Japan conf. on computation of differential equation-
sand dynamical systems, pages 1–28. World Scientific, Singapore, (1993).
[Fen95] K. Feng: Collected works of Feng Kang. volume I,II. National Defence Industry
Press, Beijing, (1995).
[FW94] K. Feng and D.L. Wang: Dynamical systems and geometric construction of algo-
rithms. In Z. C. Shi and C. C. Yang, editors, Computational Mathematics in China, Con-
temporary Mathematics of AMS, Vol 163, pages 1–32. AMS, (1994).
[Gei03] H. Geiges: Contact geometry. Math.SG/0307242, (2003).
[MNSS91] R. Mrugała, J.D. Nulton, J.C. Schon, and P. Salamon: Contact structure in thermo-
dynamic theory. Reports on Mathematical Physics, 29:109C121, (1991).
[QZ92] M. Z. Qin and W. J. Zhu: Construction of higher order symplectic schemes by com-
position. Computing, 47:309–321, (1992).
[Shu93] H.B. Shu: A new approach to generating functions for contact systems. Computers
Math. Applic., 25:101–106, (1993).
Chapter 12.
Poisson Bracket and Lie–Poisson Schemes
1. Bilinearity
2. Skew-Symmetry
{F, H} = −{H, F }.
3. Jacobi Identity
4. Leibniz Rule
Ḟ = {F, H}, ∀ F ∈ C ∞ (M )
500 12. Poisson Bracket and Lie–Poisson Schemes
is called the generalized Hamiltonian equation. The most general case of Hamiltonian
system is the one with symplectic structure, whose equations have the form:
5 6 5 6
O I p
ż = JHz , J = , z= .
I O q
According to Darboux theorem, a general Poisson system with finite dimensions
can be transformed into a local coordinate form, whose equations may be written as
ż = K(z)Hz , (1.1)
the corresponding Poisson bracket is
{F, H} = (∇z F (z))T K(z)∇z H(z), ∀ F, H ∈ C ∞ (M ).
K(z) satisfies 4 properties the above , if and only if K(z) = (kij (z)) satisfies
∂klm (z) ∂k ∂k
kij (z) + kil (z) mj + kim (z) jl = 0, j, l, m = 1, 2, · · · , n. (1.2)
∂zi ∂zi ∂zi
We remark that any antisymmetry constant matrix satisfies (1.2) and hence is a Hamil-
tonian operator, and the bracket defined by it is a Poisson bracket. We will discuss its
algorithm in more detail in the next section.
Definition 1.1. A diffeomorphism z → z = g(z) : M → M is called a Poisson
mapping, if it preserves the Poisson bracket, i.e.,
{F ◦ g, H ◦ g} = {F, H} ◦ g, ∀ F, H ∈ C ∞ (M ). (1.3)
Theorem 1.2. For a Poisson manifold with structure matrix K(z), Equation (1.3) is
equivalent to
gz K(z)gzT = K( z ),
where gz is the Jacobian matrix of g with respect to z.
Proof.
T
{F ◦ g, H ◦ g} = ∇(F ◦ g) K(z)∇ H ◦ g(z)
= (F ◦ g)z K(z)(H ◦ g)T
z
T T
∂g ∂g
= Fz (g(z)) K(z) Hz (g(z)
∂z ∂z
T
∂g ∂g
= (∇F ◦ g)T K(z) (∇H ◦ g),
∂z ∂z
and
{F, H} ◦ g = ∇F T K∇H(g(z)) = (∇F ◦ g)T K(g(z))(∇H ◦ g).
By comparison, we get
gz (z)K(z)(gz (z))T = K(g(z)) = K(
z ).
The theorem is proved.
12.1 Poisson Bracket and Lie–Poisson Systems 501
It is easy to verify that {F, H} satisfies the 4 properties of a Poisson bracket. For
the infinite dimensional evolution equations, there exists a corresponding coordinate
definition; see the literatures[Arn89,MR99] .
502 12. Poisson Bracket and Lie–Poisson Schemes
Lie group action and momentum mapping. The Lie–Poisson system is closely re-
lated to the Hamiltonian system with symmetry.
Definition 1.6. The invariant property of a Hamiltonian system under one parameter
differomorphism group is called symmetry of the Hamiltonian system. Under certain
circumstance, this invariant property is called momentum. The corresponding map-
ping is called momentum mapping.
The Lie group, action on manifold M , ∀ g ∈ G, and corresponds to a self-homeomorph-
ism φg on M . Below, we consider only the translation action of G on itself and the
induced action on T G and T ∗ G.
Definition 1.7. Infinitesimal generator vector field: let g be a Lie algebra of G, ξ ∈g,
then exp tξ ∈ G,
d
ξM = φexp tξ (x), x ∈ M
d t t=0
is called infinitesimal generator vector field of the flow Ft = φexp tξ .
∀ ξ ∈ g, d J(ξ) = iξp ω,
where J(ξ) is defined by J(ξ)(x) = J (x), ξ, · , · denotes a scalar product, and
ξp is the infinitesimal generator of the action to ξ.
Proof. See[MW83] .
φg
P −−−−→ P
⏐ ⏐
⏐ ⏐
JG JG
Ad∗
g −1
g∗ −−−−→ g∗
Corollary 1.13. Let φ be a Poisson action of G on the manifold M , and φ" be the lifted
action on T ∗ (M ) = P . Then this action φ" is symplectic and has an Ad∗ -equivariant
momentum mapping given by
Below, we will discuss the translation action of a Lie group on itself using the
above theorem and deduction.
Let G be a Lie group, φ : G × G → G be a left translation action (g, h) → gh.
Then its infinitesimal generator is
ξG (g) = Te Rg ξ = Rg∗ ξ.
Because lifted action is symplectic, by Corollary 1.13, we can obtain the momentum
mapping:
or can rewrite it as
JL (αq ) = Rg∗ αq .
Likewise, we can obtain the similar result for the right translation
JR (αq ) = L∗g αq .
Lie–Poisson bracket and motion equation. In the previous sections, we have intro-
duced the Lie–Poisson bracket and equations which are expressed by the local coordi-
nates. Below, we will introduce an intrinsic definition of Lie–Poisson bracket and its
induced equation of motion.
δF
Let · , · be the pairing between g∗ and g, ∀ F : g∗ → R, ∈ g, μ ∈ g∗ , is
δμ
defined by
= δF>
DF (μ)γ = γ, , γ ∈ g∗ .
δμ
504 12. Poisson Bracket and Lie–Poisson Schemes
where [ · , · ] is the Lie bracket on g. The above equation is usually denoted as {F, G}.
It is easy to verify that { · , · } satisfies the 4 properties of Poisson bracket, and are
often called as (−) Lie–Poisson bracket. They are first proposed by Lie[Lie88] and
are redefined by Berezin and others thereafter. We can prove that { · , · } can be derived
from the left translation reduction of a typical Poisson bracket on T ∗ G. If the right
translation reduction is used, we have the Lie–Poisson bracket (+):
= δF δG >
{F, G}(μ) = μ, , = {F, G}+ .
δμ δμ
Proof. Because
= δF >
Ḟ (μ) = DF (μ) · μ̇ = μ̇, ,
δμ
and
= δF δH > = δF > = ∗ δF >
{F, H− }− (μ) = − μ, , = μ, ad δH = ad δH μ, .
δμ δμ δμ δμ δμ δμ
Since F is arbitrary, we obtain
μ̇ = ad∗δH μ.
δμ
μ̇ = −ad∗δH μ.
δμ
Henceforth, we will denote the system of left translation reduction as g∗+ , and the right
translation reduction as g∗− . Generally speaking, the rigid body and Heavy top system
belongs to the left invariant system g∗− , and the continuous systems, such as plasma
and the incompressible flow, are right invariant system g∗+ .
Proof. See[MW83] .
12.1 Poisson Bracket and Lie–Poisson Systems 505
From this lemma, we can obtain the following reduction theorem (it will be used
in the generating function theory later).
Theorem 1.16. 1◦ ∗
For the left invariant system g− , we have the following diagram
commutes:
GtH◦J
T ∗ G −−−−−→
R
T ∗G
⏐ ⏐
⏐
JR G
⏐
JR G
∗ Gt ∗
g− −−−H−→ g−
where H : g∗ → R is a Hamiltonian function on g∗− , GtH is a phase flow of Hamilto-
nian function H on g∗− , and GtH◦JR is phase flow of Hamiltonian function H ◦ JR on
T ∗ G.
2◦ Similarly for right invariant system g∗+ , we have
GtH◦J
T ∗ G −−−−→
L
T ∗G
⏐ ⏐
⏐
JL G
⏐
JL G
Gt
g∗+ −−−H−→ g∗+
Theorem 1.17. The solutions of a Lie–Poisson system are a bundle of coadjoint or-
bits. Each coadjoint orbit is a symplectic manifold and is called symplectic leave of
the Lie–Poisson system.
This theorem is from literature[AM78] . For Lie–Poisson system such as Heavy Top
and the compressible flows, similar set of theories can be established. The readers can
refer to literature[MRW90] for more details.
Because the system that takes kinetic energy T as the Hamiltonian function is left
invariant, MS (t) is the momentum mapping exactly.
Corollary 1.19. Euler equation
d
MB (t) = {WB (t), MB (t)} = {A−1 MB (t), MB (t)}, (1.11)
dt
{ξ, a} = ad∗ξ a, ∀ ξ ∈ g, a ∈ g∗ .
Proof. 1◦ From the Lie–Poisson equation of the motion μ̇ = ad∗∂H μ, we can obtain
∂μ
directly
1 1 −1
H = (WB (t), MB (t)) = A MB (t), MB (t) ,
2 2
δH
= A−1 MB (t) = WB (t).
δ MB
2◦ By the definition of spatial angular momentum, we have
Since
MS (t) = MS (0) =⇒ Ad∗g(t)−1 MB (t) = Ad∗g(0)−1 MB (0) = η.
This also indicates that the trajectory of Lie–Poisson equation lies in some coadjoint
orbit. From
since
d
Adg(t) ξ = [T Rg(t)−1 ġ(t), Adg(t) ξ]
dt
12.2 Constructing Difference Schemes for Linear Poisson Systems 507
(see[AM78] ), then
= >
d MB (t)
,ξ = η, [T Rg(t)−1 ġ(t), Adg(t) ξ]
dt
= η, Adg(t) [T Lg(t)−1 ġ(t), ξ]
= Ad∗g(t) η, adT Lg(t)−1 ġ(t) ξ
= ad∗T L MB (t), ξ
g(t)−1 ġ(t)
d MB (t)
=⇒ = ad∗T L −1 ġ(t) MB (t) = {WB (t), MB (t)}.
dt g(t)
∂H ∂H
ġ(t) = T Lg(t) = Lg(t)∗ , (1.13)
∂μ ∂μ
μ̇(t) = ad∗∂H μ(t). (1.14)
∂μ
Its solution is μ(t) = Ad∗g(t) Ad∗g(0)−1 μ(0). The Equation (1.14) is called as the Lie–
Poisson equation.
Since the phase flow of Hamiltonian system is Poisson phase flow, which preserves
the Poisson structure, it is important to construct difference schemes for system (1.4)
that preserve the same property. Difference scheme that preserves the Poisson bracket
is called as the Poisson difference scheme.
One special case of the Poisson phase flow is the symplectic phase flow. How to
construct the symplectic difference schemes has already been described in the previ-
ous chapters. The reader can also refer to literatures[Fen85,FWQW89,FQ87,CS90] for more
details. However, the numerical algorithm for a general Poisson phase flow is still
in its infancy. So far the results are limited to cases where structure matrix K is
constant[Wan91,ZQ94,AKW93,Kar04] and K(z) is linear (Lie–Poisson) only. We will dis-
cuss the results for the Lie–Poisson case in the next section. In this section, we will
discuss the results when K is a constant matrix.
508 12. Poisson Bracket and Lie–Poisson Schemes
is an element of SK (n) .
For a generalized Cayley transformation, we have the following result similarly:
p(Λ)
Theorem 2.3. Given φ(Λ) = , p(Λ) is a polynomial that satisfies p(0) = 1,
p(−Λ)
ṗ(0) = 0, if B ∈ sK (n), then
A = φ(B) ∈ SK (n).
∞
x
ex = Ck Jk (R)Qk , |x| < R, (2.1)
R
k=0
Q0 (x) = 1,
Q1 (x) = x,
Qk+1 (x) = Qk−1 (x) + 2xQk (x),
and applying the Chebyshev spectral method to the numerator and denominator re-
spectively, we can obtain the Poisson algorithm.
It was pointed out in literature[TF85] that when k > R, the series converges ex-
ponentially. Therefore, the summation in (2.1) is always finite. Where to truncate the
series is determined by the size of Jk (R). Since Jk (R) converges exponentially too,
only a few steps of iteration is enough. Numerical tests show that this method has high
accuracy and efficiency, especially when A is a dense matrix. The above method can
be applied only to the linear dynamic system, where H is a quadratic form of z,
ż = KBz.
i.e.,
1 1
I − τ KHzz (I + B) zz = I + τ KHzz (I − B),
2 2
∂x
where z k+1
= z, z = z, xy =
k
, therefore,
∂y
−1
1 1
zz = I − τ KHzz (I + B) I + τ KHzz (I − B) .
2 2
zz K zzT = K,
i.e.,
−1
1 1
I − τ KHzz (I + B) I + τ KHzz (I − B) K
2 2
T −1
1 1
· I + τ KHz z(I − B) I − τ KHz z(I + B) = K.
2 2
τ τ
GτH,∓B = GH,−B
2
◦ GH,B.
2
Proposition 2.4. The above scheme has the second-order accuracy and the following
proposition can be easily derived.
1
If φA (z) = z T Az, where AT = A, is a conservative quantity of Hamiltonian
2
dz ∂H
system = K , and if A satisfies B T A + AB T = 0, then φA (z) is also a
dt ∂z
conservative quantity of difference scheme GτH,−B .
then
1
z + z)T A(
( z + z) = 0.
2
From B T A + AB T = 0, we obtain
1 1
z − z))T A(
(B( z − z)B T A(
z − z) = ( z − z)
2 2
1
z − z)T (B T A + AB)(
= ( z − z) = 0,
4
T
1 1
( z − z)
z + z) + B( z − z) = 0
A(
2 2
=⇒ τ wT AKHz (w) = 0, ∀ w ∈ Rn .
Let
1 1
w = (z k+1 + z k ) + B(z k+1 − z k ),
2 2
we obtain
1 k+1 T k+1 1
(z ) Az = (z k )T Az k .
2 2
The proof can be obtained.
1 2
H= (z + z22 + z32 ),
2 1
then
∂H
ż = K = Kz.
∂z
To make scheme
1 1
z k+1 = z k + τ K∇H (I + B)z k+1 + (I − B)z k
2 2
ż = J(z)Hz = f (z).
where
1 1 1 1
(
z + z) + B(
w= z − z) = (I + B) z + (I − B)z.
2 2 2 2
The Jacobian matrix of map z → z is
% &
∂
z ∂w
A= = I + τ D∗ f (w) ,
∂z ∂z
where
⎡ ⎤
I 2 − I3 I 2 − I3
0 z3 z2
⎢ I2 I3 I2 I3 ⎥
⎢ ⎥
⎢ I 3 − I1 I 3 − I1 ⎥
D∗ f (z) = D∗ J(z)Hz = ⎢
⎢ I 1 I 3 z3 0 z1 ⎥,
⎥
⎢ I1 I3 ⎥
⎣ ⎦
I 1 − I2 I 1 − I2
z2 z1 0
I1 I2 I1 I2
∂w 1 1
= (I + B)A + (I − B),
∂z 2 2
therefore
1
When B = O, the above equation is the midpoint scheme, w = ( z + z). It is
2
easy to verify that the last equality in the above equations is dissatisfied. Hence the
midpoint scheme is not a Poisson scheme. When B = O, after complex computation,
we can obtain similarly that there does not exist any B ∈ gl(n) to satisfy the above
3 formulas. Therefore, there does not exist a Poisson scheme in a generalized Euler
form.
Gt =P
g∗ −−−H−−−→ g∗
the phase flow determined by H on g∗ can induce a phase flow on T ∗ G determined
by H ◦ JR . Let ut (q, q0 ) be a first kind generating function of the symplectic map S.
Then we have the following properties.
Property 3.1. If u : G × G → R is invariant under the left action of G, i.e.,
ut (gq, gq0 ) = ut (q, q0 ), (3.1)
then the symplectic mapping generated by u, S : (q0 , p0 ) → (p, q), where:
∂ut (q, q0 ) ∂ ut (q, q0 )
p0 = − , p= , (3.2)
∂q0 ∂q
preserves momentum mapping JL . That is to say,
JL ◦ S = JL .
For the right-invariant translation on G,
JR ◦ S = JR .
12.3 Generating Function and Lie–Poisson Scheme 515
Definition 3.2. If G acts on the configuration space without fixed point, then we say
G acts on G freely.
Property 3.3. If G acts on G freely, and its induced symplectic mapping S preserves
the momentum mapping JL , then the first-kind generating function of S is left invari-
ant.
Proof. See[GM88] .
For a left-invariant system, such as a generalized rigid body, the Hamiltonian func-
tion is left invariant, the phase flow is also left invariant, the momentum mapping JL is
a first integral for this dynamics, i.e., JL is invariant under the phase flow of GtH◦JR .
Therefore, if the action is free (generally speaking, the action is locally free), the first-
kind generating function is left invariant.
Let ut (q, q0 ) be the first-kind generating function of S, then by the left invariance
ut (q, q0 ) = ut (e, q −1 q0 ) = u
"t (g), g = q −1 q0 .
V (q) = q −1 ,
V ∗ = −L∗q−1 Rq∗−1 ,
then
∂u"
p = −L∗q−1 Rq∗−1 Rq∗0 ,
∂ g g=q−1 q0
therefore,
∂u"
μ0 = L∗q0 p0 = −L∗q0 L∗q−1
∂ g g=q−1 q0
∂u" ∂u"
= −L∗q−1 q0 = −L∗g ,
∂ g g=q−1 q0 ∂ g g=q−1 q0
and
u
∂"
μ = L∗q p = −L∗q L∗q−1 Rq∗−1 Rq∗0
∂g g=q−1 q0
u
∂" u
∂"
= −Rq∗−1 q0 = −Rg∗ .
∂g g=q−1 q0 ∂g g=q−1 q0
∂u"
M0 = Rq∗0 p0 = −Rq∗0 L∗q−1 ,
∂ g g=q−1 q0
u
∂"
M = Rq∗ p = −Rq∗ L∗q−1 Rq∗−1 Rq∗0
∂g g=q−1 q0
∂u"
= −L∗q−1 Rq∗0
∂ g g=q−1 q0
∂u"
= −Rq∗0 L∗q−1 = M0 ,
∂ g g=q−1 q0
i.e.,
JL ◦ S = JL .
Take g = q −1 q0 , then
⎧
⎪ "(g)
∗∂u
⎪
⎨ μ0 = −Lq ∂ g ,
(3.3)
⎪
⎩ μ = −Rg∗ ∂ u"(g) = Ad∗g−1 μ0 ,
⎪
∂g
"t (q −1 q0 ) = u
therefore ut (q, q0 ) = u "t (g) defines a Poisson mapping:
μ0 → μ = Ad∗g−1 μ0 .
Because
∂u ∂u
pd q − p0 d q0 = dq + d q0 ,
∂q ∂ q0
∂u ∂u ∂u ∂u
du = dq + d q0 + d t = pd q − p0 d q0 + d t,
∂q ∂ q0 ∂t ∂t
we have
% &
∂u
d pd q − p0 d q0 + d t = 0.
∂t
Note that
12.3 Generating Function and Lie–Poisson Scheme 517
d (pd q − p0 d q0 ) = d p ∧ d q − d p0 ∧ d q0
% & % &
∂p ∂p ∂p ∂q ∂q ∂q
= d p0 + d q0 + dt ∧ dp0 + dq0 + d t − d p0 ∧ d q 0
∂ p0 ∂q0 ∂t ∂p0 ∂ q0 ∂t
% &
∂p ∂q ∂p ∂q
= − d p 0 ∧ d q 0 − d p0 ∧ d q 0
∂ p0 ∂ q0 ∂ q0 ∂ p0
∂q ∂p ∂q ∂ p0
+ d p0 ∧ d t + d q0 ∧ d t
∂ t ∂p0 ∂t ∂ q0
∂p ∂q ∂p ∂p
− dp0 ∧ dt − d q0 ∧ d t
∂ t ∂p0 ∂ t ∂ q0
= f1 + f2 + f3 .
Since (p0 , q0 ) → (p, q) is symplectic, we have
gz JgzT = J =⇒ f1 = 0.
Because
⎧ ⎧
⎪
⎪
∂q ∂H ⎪
⎪
∂H
⎨ ∂t = ∂p ⎨ f2 = ∂ p d p ∧ d t
=⇒ ,
⎪
⎪ ⎪
⎪
⎩ ∂ p = −∂ H ⎩ f3 = ∂ H d q ∧ d t
∂t ∂q ∂q
therefore, d p ∧ d q − d p0 ∧ d q0
∂H ∂H
= dp ∧ dt + dq ∧ dt = dH ∧ dt
∂p ∂q
% &
∂H
=⇒ d H ∧ d t + d ∧ d t = 0.
∂t
We have % &
∂H
d H+ ∧ d t = 0.
∂t
Therefore,
∂u
+ H(p, q, t) = c.
∂t
Taking a proper initial value, we can obtain:
∂u
+ H(p, q, t) = 0,
∂t
i.e.,
∂ ut (p, q)
+ H ◦ JR (p, q, t) = 0.
∂t
Therefore we obtain the LPHJ equations
∂ u(g) ∂ u(g)
+ H − Rg∗ = 0, (3.5)
∂t ∂g
g = q −1 q0 .
518 12. Poisson Bracket and Lie–Poisson Schemes
Remark 3.4. If we can construct a generating function u(g), we then have u(q0 , q).
This function can generate a symplectic mapping on T ∗ G. By the commutative dia-
gram, a Poisson mapping on g∗ can also be induced. This is a key point of constructing
a Lie–Poisson integrator by generating function.
Remark 3.5. In order that the induced phase flow is a Poisson phase flow, the phase
flow on T ∗ G should be symplectic. Therefore, the condition of g = q −1 q0 cannot be
discarded. Namely, when t → 0, g = q −1 q0 (unit element).
To make sure
JL (p0 , q0 ) = JL (p, q) =⇒ Ad∗q−1 JR (p0 , q0 ) = Ad∗q−1 JR (p, q)
0
Remark 3.7. The above generating function theory can be transformed into the gen-
erating function theory on g (for details see literature[CS90] ). That is to say, the above
generating function theory on T ∗ G can be reformulated by the exponential mapping in
terms of algebra variables, which has been done by Channell and Scovel[CS90] . Below,
we list only some of their results.
For g ∈ G, choose ξ ∈ g, so that g = exp (ξ). Then the LPHJ equation can be
transformed into ⎧ ∂s
⎪
⎪ + H(−ds · ψ(adξ )) = 0,
⎪
⎨ ∂t
M0 = −ds · χ(adξ ), (3.6)
⎪
⎪
⎪
⎩
M = −ds · ψ(adξ ),
where ⎧
⎨ χ(adξ ) = id + 1 adξ + 1 ad2ξ + · · · ,
2 12 (3.7)
⎩ −adξ
ψ(adξ ) = χ(adξ ) · e χ(adξ ) − adξ ,
−1
and the condition g = q q0 is transformed into
i.e.,
ξ|t=0 = id.
12.3 Generating Function and Lie–Poisson Scheme 519
= 0.
and then substitute it into Equation (3.6). Next, we calculate M = exp (ξ)M0 . On
repeating this procedure, we can obtain a Lie–Poisson algorithm.
Below, we will apply this algorithm to free rigid body. For motion of the rigid
body, χ(ξ) has a closed expression (see Subsection 12.5.2). Solving nonlinear (3.12)
for ξ becomes a key point. It is necessary to linearize (3.12). The iterative formula for
ξ is
1 + τ c1 ξ − c3 ξ(ξ + c4 ) (I −1
M0 × ξ) + c2 (I
−1 p ) δ ξ = ξ
0 k+1 − ξk ,
where
2 − |ξ| sin |ξ| − 2 cos |ξ| cos |ξ| − 1
c1 = , c2 = ,
|ξ|4 |ξ|2
−2|ξ| − |ξ| cos |ξ| + 3 sin |ξ| 2|ξ| − sin |ξ|
c3 = , c4 = .
|ξ|5 |ξ|3
(ξ, ξ)
where u0 = generates the identical transformation at time t = 0. Substituting
2
(3.13) into the LPHJ equation, we have
∂H
u1 = −H(V ), u2 = · du1 · ψ(adξ ), · · · . (3.14)
∂V
Below, we will take so(3)∗ as an example to explain the flaw of this algorithm.
ξ2
For so(3)∗ , u0 = , and hence V = ξ. The first-order scheme is
2
ξ2 ξ2 τ
S1 = u0 + τ u1 = − τ H(ξ) = − ξI −1 ξ.
2 2 2
τ2 ξ2 τ 2 ∂H
S2 = S 1 + u2 = − τ H(ξ) + · du1 · ψ(adξ )
2 2 2 ∂V
ξ2 τ τ2
= − ξI −1 ξ − I −1 ξ I −1 ξ · ψ(ξ) .
2 2 2
Using the system of Equation (3.6) (for SO(3) M, M0 denote angular momentum),
we get:
M − M0 = −du · adξ . (3.15)
Next, we will prove that S1 indeed generates a first-order Lie–Poisson scheme to
the Euler equation. However, S2 actually is not a second-order approximation to the
Euler equation. Furthermore, we will find that with this algorithm, it is impossible to
construct difference scheme that preserves the momentum mapping.
Because % 2 &
ξ τ
d S1 = d − ξ · I −1 ξ = ξ − τ I −1 ξ
2 2
M − M0 = (ξ − τ I −1 ξ) · adξ = −τ I −1 ξ · adξ
= τ [ξ, I −1 ξ] = τ [−M0 + O(τ ), I −1 (−M0 + O(τ ))]
= τ [M0 , I −1 M0 ] + O(τ 2 ).
Ṁ = [M, I −1 M ]. (3.16)
For the second-order generating function S2 , we first calculate χ(ξ). Let χ(ξ) =
1+a1 ξ +a2 ξ 2 , where a1 , a2 have closed analytical expression (see Subsection 12.5.2)
as follows
1 − cos |ξ|
a1 = 2 ,
sin |ξ| + (1 − cos |ξ|)
2
therefore,
u2 = −I −1 ξ(I −1 ξ · ψ(ξ))
= −I −1 ξ, I −1 ξ − a2 I −1 ξ(I −1 ξ · ξ 2 ),
then
τ2
d S2 = ξ − τ I −1 ξ − τ 2 (I −1 )2 ξ − d a2 I −1 ξ · (I −1 ξ · ξ 2 ) ,
2
by
522 12. Poisson Bracket and Lie–Poisson Schemes
M − M0 = −dS2 · adξ
τ2
= −(ξ − τ I −1 ξ − τ 2 (I−1 ξ)2 − d(a2 I −1 ξ · (I −1 ξ · ξ2 ))) · ξ
2
= τ [M0 , I −1 M0 ] + a1 τ 2 [[M0 , I −1 M0 ], I −1 M0 ]
+[M0 , I −1 [M0 , I −1 M0 ]]
+a2 I −1 M0 (I −1 M0 · M 42 ) + I −1 (I −1 M0 · M 42 )M0
0 0
τ2
− d a2 · I −1 ξ(I −1 ξ · ξ2 ) · ξ + O(τ 3 ).
2
(3.17)
According to the Euler equation (3.16), its second-order approximation should be
τ2
M − M0 = τ [M0 , I −1 M0 ] + ([[M0 , I −1 M0 ], I −1 M0 ]
2 (3.18)
+[M0 , I −1 [I −1 M0 , M0 ]]) + O(τ 3 ).
Let Λ(t) ∈ SO(3), such that Λ(t)Λ(t)T = I, |Λ(t)| = 1. Then the equation of motion
for the free rigid body can be formulated as
4 (t),
Λ̇(t) = Λ(t)W (4.1)
4 (t) ∈ so(3), so(3) is the Lie algebra of SO(3). The isomorphism relation,
where W
so(3) R3 , can be realized through the following equations:
4 (t) W (t) ∈ R3 ,
W
⎡ ⎤
⎡ ⎤ w1
0 −w3 w2
⎢ ⎥
⎢ ⎥ ⎢ w ⎥
⎢ w3 0 −w1 ⎥ ⎢ 2 ⎥,
⎣ ⎦ ⎢ ⎥
⎣ ⎦
−w2 w1 0 w3
4 (t) · a = W × a,
W a ∈ R3 .
where J is called inertia operator, M the body angular momentum. The body variables
and the spatial variables have the following relations:
⎧
⎪ ω = AW,
⎪
⎨
m = ΛM, ω 4 ΛT =⇒ ω = ΛW,
= ΛW
⎪
⎪
⎩
a = ΛA,
here A is an acceleration.
Operator “ ” has the following equalities:
524 12. Poisson Bracket and Lie–Poisson Schemes
u
× v = [
u, v],
· v = u × v,
u
u, v] · w = (u × v) × w,
[
1
u·v = u v).
tr (
2
The equation of motion of the rigid body may be expressed on space SU (2) or SH1
(unit quaternion). Applying their equivalence (their Lie algebra is isomorphism), we
may obtain different forms of the Equation (4.1) under SU (2) and SH1 .
SU (2): U ∈ SU (2), satisfies
U U ∗ = I, |U | = 1.
U̇ = U Ωu ,
as a basis, where
5 6 5 6
0 1 0 −i
σ1 = , σ2 = ,
1 0 i 0
5 6 5 6
1 0 1 0
σ3 = , σ0 =
0 −1 0 1
are 4 Pauli matrices.
It is easy to see that
5 6
3
−iω3 −ω2 − iω1
ωi σi = ∈ SU (2).
i=1 ω2 − iω1 iω3
Hence
Ωu = (ω1 , ω2 , ω3 ) ∈ su(2) R3 so(3),
using the matrix notation, rewrite the equation:
5 6 5 65 6
σ̇ β̇ σ β −iω3 −ω2 − iω1
= .
γ̇ δ̇ γ δ ω2 − iω1 iω3
Ωh = ω1 i + ω2 j + ω3 k = (0, ω1 , ω2 , ω3 ), ωh = (ω1 , ω2 , ω3 ).
After the above transformation, the equation of motion becomes more simpler. The
number of unknowns become fewer from the original 9 (SO(3)) to 4 complex vari-
ables (SU (2)), and then reduced to 4 real variables (SH1 ). The computation storage
and operation may be sharply reduced for large-scale scientific computations.
More details about the relations among SO(3), SU (2) and SH1 will be given in
Section 12.5.
The total energy and the angular momentum, especially the angular momentum,
are important invariants for the rigid motion. Many experiments indicated that the
energy and the angular momentum can be well maintained, which is essential for
computer simulation to have a good approximation to the real motion.
The equation of motion for the rigid body is
⎧
⎨ Λ̇(t) = Λ(t)W 4 (t),
/
Ṁ (t) = M (t) × W (t)
⎩ =⇒ I · Ẇ (t) = IW (t) × W (t),
M (t) = I · W (t)
Mn · I −1 Mn+1 = Mn+1 · I −1 Mn .
= Λ−1
n+1 Λn Mn .
12.4 Construction of Structure Preserving Schemes for Rigid Body 527
Λn+1 Mn+1 = Λn Mn .
By comparison, we obtain
δt −1 δt
Λn+1 = Λn I − I −1 (Mn+1 + Mn )
I + I −1 (Mn+1 + Mn ) .
4 4
Since
δt −1 δt
I (Mn+1 + Mn ) = Wn + O(δt2 ),
4 2
from Cayley transformation, we know this is a second-order approximation to equa-
tion Λ̇ = ΛW4.
In brief, if we construct an energy-preserving scheme on so(3)∗ , we may obtain
a scheme approximate to the equation of motion by using the conservation of an an-
gular momentum. We remark that this highly depends on the schemes constructed on
so(3)∗ . Not every scheme on so(3)∗ corresponds to a good approximation scheme to
the equation of motion on SO(3). Ge–Marsden algorithm for Lie–Poisson system is
a typical example.
The orbit-preserving[LQ95a] here means the motion trajectory remains at coadjoint or-
bit. For rigid body this means in every time step
Mn+1 = Λn Mn , ∃ Λn ∈ SO(3).
4 · M,
Ṁ = M × W = −W × M = −W 4 ∈ SO(3),
W 4 = I −1 M.
W
(3)
4n Mn + M̈ δt2 + M
Mn+1 = Mn − δtW δt3 + · · ·
2 3!
2
4n Mn + δt
= Mn − δtW (Mn × Wn × Wn ) (4.6)
2
δt2
+ Mn × I −1 (Mn × Wn ) + · · · .
2
Let
b(δt) = δtB1 + δt2 B2 + δt3 B3 + · · · ,
substitute it into (4.5), and retain only the first two terms
1
Mn+1 = Mn + δtB1 Mn + δt2 B2 Mn + (δtB1 + δt2 B2 )2 Mn + o(δt3 )
2
1
= Mn + δtB1 Mn + δt2 B2 Mn + δt2 B12 Mn + o(δt3 ).
2
(4.7)
Comparing the coefficients of Equation (4.6) with those of (4.7), we have
4n ,
B1 = −W
(B12 + 2B2 )Mn = (Mn × Wn × Wn ) + (Mn × I −1 (Mn × Wn ))
4 2 Mn − I −1 (M
= W n × W n ) Mn ,
n
then
4n ,
B1 = −W
1 −1
B2 = − I (M
n × Wn ) .
2
Likewise, we can construct third or fourth order schemes. Here we give only the result
1 4 −1 4 + I −1 (M ×
B3 = W I (M × W ) + 2I −1 (M
× W )W W × W)
6
1
+ I −1 M × I −1
1
(M × W ) − B1 B2 − B2 B1 .
2 2
Another way to construct the orbit-preserving scheme is the modified R–K method,
which can be described as follows.
If the initial value M0 is known, let:
12.4 Construction of Structure Preserving Schemes for Rigid Body 529
μ0 = M0 ,
−1 μ )
μ1 = eτ c10 (−I 0
M0 ,
−1 μ ) τ c (−I
−1 μ )
μ2 = eτ c21 (−I 1
e20 0
M0 ,
···
−1
−1
−1 μ )
μr = eτ cr,r−1 (−I μr−1 )
eτ cr,r−2 (−I μr−2 )
· · · eτ cr,0 (−I 0
M0 ,
τ2 2 τ2
+ c0 (I −1 M0 )2 + c21 (I
−1 M )2 + τ 2 c c (I
0 0 1
−1 M )2 + O(τ 3 ) M
0 0
2 2
M0 · M0 + τ 2 c1 c10 I −1 (M
= M0 − τ (c0 + c1 )I −1 0×I
−1 M ) · M
0 0
τ2 2
+ (c + c21 + 2c0 c1 )(I
−1 M )2 M + O(τ 3 ).
0 0
2 0
1 1
Set c0 = c1 = , c10 = 1 or c0 = 0, c1 = 1, c10 = , we obtain a second-order
2 2
modified R–K method.
Literature[CG93] gives the modified R–K methods for general dynamic system. The
scheme on so(3)∗ constructed via the above methods can be written as Mn+1 = ΛMn .
Take Λ−1
n+1 Λn = Λ, we obtain Λn+1 = Λn Λ
−1
. It is easy to verify that the Λn+1 =
−1
Λn Λ approximates Λ̇ = ΛW in the same order of accuracy as scheme Mn+1 =
ΛMn .
1 1 1
H = H − a1 |x|2 = (a2 − a1 )x22 + (a3 − a1 )x23 = H1 + H2 ,
2 2 2
1 1
where H1 = (a2 − a1 )x22 , H2 = (a3 − a1 )x23 .
2 2
12.4 Construction of Structure Preserving Schemes for Rigid Body 531
where
⎡ ⎤
0 −x3 x2
⎢ ⎥
J(x) = ⎢
⎣ x3 0 −x1 ⎥
⎦.
−x2 x1 0
where x2 is a constant.
Among symplectic difference schemes for the standard symplectic system (4.10),
only a few of them can preserve the Lie-Poisson structure of the original system (4.9).
Theorem 4.1. For the system (4.9), the midpoint scheme is a Lie–Poisson scheme[LQ95a] .
In order to prove the Theorem 4.1, we need the following lemma first.
Lemma 4.2. For the system (4.9), a symplectic algorithm for the standard symplectic
system (4.10) preserves Poisson structure, if and only if the following three conditions
are satisfied
⎧
⎪ −x11 x3 + x13 x1 = − x3 ,
⎪
⎨
x31 x3 − x33 x1 = − x1 , (4.11)
⎪
⎪
⎩
1 + x32 x
x12 x 3 = 0,
∂ xi
where xi = xni , xi = xn+1
i , xij = .
∂ xj
⎡ ⎤⎡ ⎤⎡ ⎤
x11 x12 x13 0 −x3 x2 x11 0 x31
⎢ ⎥⎢ ⎥⎢ ⎥
⎢ 0 1 0 ⎥ ⎢ −x1 ⎥ ⎢ x32 ⎥
⎣ ⎦ ⎣ x3 0 ⎦ ⎣ x12 1 ⎦
x31 x32 x33 −x2 x1 0 x13 0 x33
⎡ ⎤
0 −
x3 2
x
⎢ ⎥
= ⎢ 3
⎣ x 0 x1 ⎥
− ⎦,
−
x2 1
x 0
i.e.,
⎡ ⎤ ⎡ ⎤
0 −x11 x3 + x13 x1 a13 0 −
x3 2
x
⎢ ⎥ ⎢ ⎥
⎢ x11 x3 − x13 x1 0 x31 x3 − x33 x1 ⎥ ⎢ 3 x1 ⎥
−
⎣ ⎦=⎣ x 0 ⎦,
−a13 x33 x1 − x31 x3 0 −
x2 1
x 0
where a13 = (x12 x3 − x13 x2 )x31 + (x13 x1 − x11 x3 )x32 + (x11 x2 − x12 x1 )x33 .
Since the scheme is symplectic for (4.10), we have
−x13 x31 + x11 x33 = 1.
So a13 can be simplified as:
a13 = (x3 x31 − x1 x33 )x12 + (x13 x1 − x11 x3 )x32 + x2 .
Comparing the corresponding elements of the matrix on both sides and using the con-
2 = x2 , we have
dition x
⎧
⎪ x11 x3 − x13 x1 = x3 ,
⎪
⎨
x31 x3 − x33 x1 = − x1 ,
⎪
⎪
⎩
1 + x32 x
x12 x 3 = 0.
Thus before the lemma is proved.
Now we will prove the Theorem 4.1.
Proof. The midpoint scheme for system (4.9) is (here, I = (I1 , I2 , I3 ) = (a1 , a2 , a3 ))
⎧
⎪ + x3
x
⎪
⎪ x1 = x1 + τ (I1 − I2 ) 3 x2 ,
⎨ 2
x2 = x2 ,
⎪
⎪
⎪
⎩ x + x1
x
3 = x3 + τ (I2 − I1 ) 1 x2 .
2
⎡ ⎤
x11 x12 x13
⎢ ⎥
Its Jacobian matrix is ⎣ 0 1 0 ⎦, where
x31 x32 x33
12.4 Construction of Structure Preserving Schemes for Rigid Body 533
⎧ τ
⎪
⎪ x11 = 1 + (I1 − I2 )x2 x31 ,
⎪
⎪ 2
⎪
⎪
⎪
⎪ τ τ
⎪
⎪ x12 = (I1 − I2 )(
x3 + x3 ) + (I1 − I2 )x2 x32 ,
⎪
⎪ 2 2
⎪
⎪
⎪
⎪ τ τ
⎨ x13 = (I1 − I2 )x33 x2 + (I1 − I2 )x2 ,
2 2
⎪
⎪ τ τ
x31 = (I2 − I1 )x11 x2 + (I2 − I1 )x2 ,
⎪
⎪
⎪
⎪ 2 2
⎪
⎪
⎪
⎪ τ
x32 = (I2 − I1 )(
τ
x1 + x1 ) + (I2 − I1 )x2 ,
⎪
⎪
⎪
⎪ 2 2
⎪
⎪
⎩ x = 1 + τ (I − I )x x .
33 2 1 2 13
2
Solving the above equations, we get
⎧
⎪ 1 − a2
⎪
⎪ x11 = x33 = ,
⎪
⎪ 1 + a2
⎪
⎪
⎪
⎪ τ
⎪
⎪ (I1 − I2 )
x3
⎪
⎪
⎨ x12 = 2 2
,
1+a
(4.12)
⎪
⎪ 2a
⎪
⎪ x13 = −x31 = − ,
⎪
⎪ 1 + a2
⎪
⎪
⎪
⎪ τ
⎪
⎪ (I2 − I1 )
x1
⎪
⎩ x32 = 2
2
,
1+a
where
τ
a= (I2 − I1 )x2 . (4.13)
2
Substituting the system of Equations (4.12) into condition (4.11), we find that all con-
ditions are satisfied. Therefore, by Lemma 4.2, the scheme is of Poisson.
Lemma 4.3. [FQ91] Consider dynamic system ẋ = a(x). If a can be split into a =
a1 + a2 + · · · + ak , and g s esa is phase flow of a dynamic system, then
s s s s
gis esai , 2nd-order, ∀ i =⇒ g12 ◦ · · · gk2 ◦ gk2 ◦ · · · g12 esa , 2nd-order.
Proof. For the standard symplectic system (4.10), the generalized Euler scheme
= x + τ J∇H(B
x x + (I − B)x)
is symplectic, iff
1
B= (I + C), JC + C T J = O. (4.14)
2
It is natural to ask what kind of symplectic difference scheme for the system (4.10)
is also a Poisson scheme for the system (4.9). Below we restrict our discussion to the
generalized Euler scheme ! (4.14).
c1 c2
Let C = , then the symplectic condition (4.14) turns into c4 = −c1 .
c3 c4
Therefore,
534 12. Poisson Bracket and Lie–Poisson Schemes
5 6
1 1 + c1 c2
B= ,
2 c3 1 − c1
then
5 6
1 x1 + (1 − c1 )x1 + c2 (
(1 + c1 ) x3 − x3 )
x + (I − B)x =
B
2 x1 − x1 ) + (1 − c1 )
c3 ( x3 + (1 − c1 )x3
5 6 (4.15)
1 z1
= ,
2 z3
x11 x3 − x13 x1 = x
3 (see (4.11)).
c1 = 0, c2 = −c3 . (4.17)
Substituting Equation (4.17) into (4.16), and recalculating the Jacobian matrix, we
have
2a(1 − ac2 ) (1 − ac2 )2 − a2
x31 = 2 , x33 = ,
a + (1 − ac2 )
2 2 a + (1 − ac2 )
2
x31 x3 − x33 x1 = −
x1
is satisfied. Likewise, we can prove that another condition of (4.11) is also satisfied.
From (4.17), we have C = cJ, where c is an arbitrary constant and
12.4 Construction of Structure Preserving Schemes for Rigid Body 535
5 6
0 1
J= .
−1 0
where
(i, j, k), i, j, k is not the same,
εijk =
0, i, j, k is the same.
There are two independent Casimir functions for bracket (4.18)
3
3
f1 = p2i , f2 = pi xi .
i=1 i=1
where square bracket denotes cross product. H is the system’s energy, x and p are
angular momentum, and momentum under momentum coordinate. For a general case,
energy H is of quadratic form about x, p, and positive definite, which can be given as
follows
536 12. Poisson Bracket and Lie–Poisson Schemes
3
3
3
2H = ai x2i + bij (pi xj + xi pj ) + cij pi pj . (4.20)
i=1 i,j=1 i,j=1
For heavy top, the energy is often expressed as the sum of kinetic energy and
potential energy, i.e.,
x21 x2 x2
H(x, p) = + 2 + 3 + γ1 p1 + γ2 p2 + γ3 p3 , (4.21)
2I1 2I2 2I3
where Ii is the main movement inertia of the rigid body, γi (i = 1, 2, 3) are three
coordinates of the center of mass. It is easy to see that this is a separable system.
The structure matrix of the Lie–Poisson system is
5 6
J(x) J(p)
,
J(p) O
x2i
Hi = , Hi+3 = γi pi , i = 1, 2, 3.
2Ii
Theorem 4.4. The Midpoint scheme of (4.22) is Poisson scheme for heavy top.
Proof. By Theorem 1.2, the midpoint scheme is the Poisson scheme iff mapping
(x, p) −→ (
x, p)
satisfies
⎡ ⎤
∂x
∂x 5
⎡
6 ∂ x ∂ p ⎤ 5 6
⎢ ∂x ∂p ⎥ J(x) J(p) ⎢ ∂ x ∂x ⎥ J(
x) J(
p)
⎢ ⎥ ⎣ ⎦= . (4.23)
⎣ ∂ p ∂ p ⎦ ∂x ∂ p
J(p) O J(
p) O
∂x ∂p ∂p ∂p
∂ y
Denote = yz , then the expand Equation (4.23),
∂z
⎧
⎪ x xT
x J(x) x = J(
x),
⎪
⎨
x pT
x J(x)x +x pT
x J(p) p = J(
p), (4.24)
⎪
⎪
⎩
pT
px J(x)x +p pT
p J(p)x +p pT
x J(p)p = 0.
From the results of Subsection12.4.4, the first equation of system (4.24) is obviously
hold. Note also
⎡ ⎤ ⎡ ⎤
0 p21 p31 1 0 0
⎢ ⎥ ⎢ ⎥
px = ⎢
⎣ 0 0 0 ⎥ ⎢
⎦ , pp = ⎣ 0 p22 p23 ⎦ ,
⎥
0 0 0 0 p32 p33
1 1
H = Hi + Hij = ai x2i + aij (xi + xj )2 ,
2 2
i.e., we can eliminate the mixed items and transform it into a sum of squares. Next,
we can construct Lie–Poisson scheme for system with Hij as Hamiltonian function.
Take H12 as an example
∂ H12
ẋ = J(x) . (4.26)
∂x
It is easy to see that x1 + x2 is a Casimir function of system. Expanding Equation
(4.26) yields
⎡ ⎤ ⎡ ⎤
ẋ1 −a12 x3 (x1 + x2 )
⎢ ⎥ ⎢ ⎥
⎢ ẋ2 ⎥ = ⎢ a12 x3 (x1 + x2 ) ⎥ . (4.27)
⎣ ⎦ ⎣ ⎦
ẋ3 a12 (x21 − x22 )
Since x1 + x2 is a constant, denote c = x1 + x2 , the Equation (4.27) becomes
⎧
⎪ ẋ1 = −ca12 x3 ,
⎪
⎨
ẋ2 = ca12 x3 , (4.28)
⎪
⎪
⎩
ẋ3 = ca12 (c − 2x2 ).
The midpoint scheme to the above equations is no longer Lie–Poisson scheme. How-
ever, we can solve system of the Equations (4.28) analytically without difficulty.
φ(q) = q T q − 1, ∀ q ∈ M.
where p− = ψ(p, q). It is easy to verify that this is an isomorphic mapping and pre-
serving the symplectic structure.
There exist constrained equations of dynamic system on CM ,
q̇ = ∂p H,
(4.29)
ṗ = −∂q H + dφ · μ.
If it is easy to construct structure-preserving scheme for the Equation (4.29) (e.g. when
(4.29) is a separable system), then we can use map ψ to induce the algorithm on T N .
Take SO(n) as an example.
1
On T M we have a Lagrangian function L(q, q̇) = tr (q̇J q̇ T ). Using the Legen-
2
1
dre transformation, we can obtain Hamiltonian function H(p, q) = tr (pJ −1 pT ) on
∗
2
T M . Therefore, using (4.29), we can obtain the constrained Hamiltonian equation of
the dynamic system: ⎧
⎨ q̇ = 1 pJ −1 ,
2 (4.30)
⎩
ṗ = dψ · μ = 2qμ,
which is a separable Hamilton system obviously. It is easy to construct the ex-
plicit symplectic difference scheme. But on T N , the Hamiltonian function becomes
1
H(p, q) = tr (I −1 (q T p)(q T p)T ), and its Hamiltonian equations are
4
⎧
⎪
⎪ −1 T ∂H
⎨ q̇ = qI (q p) = ∂ p ,
(4.31)
⎪
⎪
⎩ ṗ = pI −1 (q T p) = − ∂ H ,
∂q
expansion system is separable. Otherwise this algorithm is impractical. Take the rigid
body as an example. On T ∗ SO(3), if Euler equation is to be solved, there are only
6 unknowns. If we expand it to T ∗ GL(n), the number of unknown becomes 18. If
system is not separable, then the computation cost will definitely increase.
2. Veselov–Moser algorithm
Veselov–Moser algorithm[MV91] is to discretize Lagrange function first and then
apply Legendre transformation to the discrete Lagrange function. The constructed al-
gorithm preserves discreted symplectic structure, thus also preserves system’s Lie–
Poisson structure. The concrete procedure is as follows:
1◦ First discretize the Lagrange function.
2◦ Add constraint and find the solution for δS = 0.
3◦ Obtain the discrete equation.
4◦ Solve this equation.
n
T
For SO(n), S = tr (Xk JXk+1 ). The constrained Lagrange function is
k=1
n
L=S+ (Xk XkT − 1),
k=1
then
δL = 0 =⇒ Xk+1 J + Xk−1 J = Λk Xk , ∀ k ∈ Z,
from this, we can have a system of equations
Mk+1 = wk Mk wk−1 ,
wk ∈ O(n), (4.32)
Mk = wkT J − Jwk ,
−1
where wk = Xk+1 Xk . It is easy to prove that this discrete system of equations con-
verges to continuous system of Euler-Arnold equations:
Ṁ = [M, Ω],
Ω ∈ o(n). (4.33)
M = JΩ + ΩJ,
To solve Equation (4.32), the key lies in solving for wk . In order to make iteration
(Xk , Yk ) → (Xk+1 , Yk+1 ) symplectic, Yk = Xk+1 , we need
This is because
T
Yk+1 J + Xk J = Xk+1 wk+1 J + Xk J
T T
= Xk+1 (wk+1 J + Xk+1 Xk J)
T
= Xk+1 (wk+1 J + wk J).
12.4 Construction of Structure Preserving Schemes for Rigid Body 541
See also
JwkT − wk J = Mk+1 = wk+1
T
J − Jwk+1 ,
then
JwkT + Jwk+1 = wk+1
T
J + wk J,
T
i.e., wk+1 J + wk J is symmetric. Thus ∃ Λk , Λk = ΛT
k , so that
T
Xk+1 (wk+1 J + wk J) = Λk Xk+1 .
Therefore,
Yk+1 J + Xk J = Λk Xk+1
satisfies symplectic condition.
The next question is how to solve wkT J − Jwk = Mk = tmk for wk ? The nu-
merical experiments show that not all solutions wk that satisfy Equations (4.32) are
the solutions we want. To solve ωk quickly, we propose to use the Quaternion method.
w ∈ SO(3) corresponds to an element q = (q0 , q1 , q2 , q3 ) in SH1 . Their relations
will be given in Section 12.5. Then the second equation in Equation (4.32) becomes
⎧
⎪
⎪ 2(α2 − α1 )q2 q1 + 2(α1 + α2 )q3 q0 = −δtm3 ,
⎪
⎨
2(α3 − α1 )q3 q1 − 2(α3 + α1 )q2 q0 = δtm2 ,
⎪
⎪
⎪
⎩ 2(α − α )q q + 2(α + α )q q = −δtm ,
3 2 3 2 3 2 1 0 1
in addition,
q02 + q12 + q22 + q32 = 1.
Solving the above nonlinear equations for (q0 , q1 , q2 , q3 ) is not an easy task. We found
that when iteration step size is small, q0 , q1 , q2 , q3 behaves reasonable. However, when
the step size is large, the solution behaves erratically. Numerical experiments show that
solving these nonlinear equations is quite time-consuming, and hence this method is
not recommended in practice.
3. Reduction method
Reduction method bases on the momentum mapping discussed in previous sec-
tions. We have mentioned that the solution of a Lie–Poisson system lies in a coadjoint
orbit in Section 12.2, and this orbit has non-degenerated symplectic structure. If we
can construct the symplectic algorithm on this reduced orbit, then this algorithm is
naturally Lie–Poisson. Moreover it preserves the Casimir function and also preserves
the orbit. Below, we will take SO(3) as an example to illustrate this method.
The coadjoint orbit of SO(3) is a two dimensional spherical surface S2r . On S2r ,
we have a symplectic structure
where x, y, z are three angular momentums in the body description. Using Euler angle
coordinate θ, ψ to do the following coordinate transformation:
⎧
⎪ x = r sin θ cos ϕ,
⎪
⎨
y = r sin θ sin ϕ,
⎪
⎪
⎩
z = r cos θ.
where % &
1 r2 sin2 θ cos2 ϕ r2 sin2 θ sin2 ϕ r2 cos2 θ
H= + + .
2 I1 I2 I3
then
⎧ dx 1 ∂H
1
⎪
⎨ dt = r ∂ x ,
2
⎪
⎩ d x2 = − 1 ∂ H .
dt r ∂ x1
This is a Hamiltonian system with standard symplectic structure, and its symplectic
algorithm is easy to construct.
To sum up, constructing Lie–Poisson scheme for a Lie–Poisson system has three
methods. The first method is to lift it to T ∗ G and construct the symplectic algorithm
(includes constraint Hamiltonian method) on it. The second is the direct construction
based on g∗ (generating function method and composition method). The third is to
construct symplectic algorithm on the reduced coadjoint orbit.
12.5 Relation Among Some Special Group and Its Lie Algebra 543
ξ ∈ so(3) =⇒ ξ + ξT = 0,
q ∈ SH1 is a normal Quaternion q = (q0 , q) = (q0 , q1 , q2 , q3 ), q = (q1 , q2 , q3 ),
We assume
⎡ ⎤
0 −ξ3 ξ2
⎢ ⎥
∀ ξ ∈ R3 , ξ = (ξ1 , ξ2 , ξ3 ) =⇒ ξ = ⎢
⎣ ξ3 0 −ξ1 ⎥
⎦ ∈ so(3),
−ξ2 ξ1 0
When A
ξ is called the axial quantity of ξ. ∈ so(3), A expresses its axial quantity.
1√
q0 = 1 + tr Λ,
2
(Q − Q23 )
⎫
q1 = 32 ⎪
⎪
4q0 ⎪
⎪
⎪
⎬
(Q13 − Q31 ) 1
q2 = =⇒ q = (Λ − ΛT ).
4q0 ⎪
⎪ 4q0
⎪
(Q21 − Q12 ) ⎪
⎪
⎭
q3 =
4q0
According to the properties of SO(3), this expansion has a closed form, i.e., the Ro-
drigue formula
% &
1
sin2 ξ 2
= 1 + sin ξ ξ + 1 % 2 & ξ .
Λ = exp (ξ) 2
ξ 2 1
ξ
2
We have two proofs of the above formula: one is from the geometry point of view and
the other is from algebra point of view. Below, we will give details on the algebraic
proof.
∀ ξ ∈ so(3), the following results hold after simple calculations:
3 4 2
ξ = −ξ,
ξ = −ξ , |ξ| = 1.
= 1 + sin ξ n2
n + (1 − cos ξ)
2 1
sin ξ 1 sin 2 ξ
2
= 1+ ξ+ % &2 ξ .
ξ 2 1
ξ
2
We can prove that ξ is the angle of rotation exp (ξ).
12.5 Relation Among Some Special Group and Its Lie Algebra 545
χ(ξ)iex(−ξ) = Idξ .
% &3
ξ ξ
For ξ ∈ so(3), from =− , we have
ξ ξ
∞
∞
∞
n
(−ξ) 2k
(ξ) 2k+1
(−ξ)
=
iex(−ξ) = +
(n + 1)! (2k + 1)! (2k + 2)!
n=0 k=0 k=0
∞
% &2 ∞
(−1)k+1 ξ2k ξ (−1)k+1 ξ2k+1 ξ
= 1+ + ·
(2k + 1)! ξ (2k + 2)! ξ
k=1 k=0
2
|ξ| − sin |ξ| cos |ξ| − 1
= 1+ ξ + ξ
|ξ|3 |ξ|2
2
= 1 + c1 ξ + c2 ξ ,
546 12. Poisson Bracket and Lie–Poisson Schemes
therefore
a1 + c1 − (a1 c2 + a2 c1 )|ξ|2 = 0,
a1 c1 + c2 + a2 − a2 c2 |ξ|2 = 0.
Solving the above equations, we have
−c1 1 − cos |ξ|
a1 = = ,
(1 − c2 |ξ|2 )2 + c21 |ξ|2 (sin |ξ|)2 + (1 − cos |ξ|)2
(cos |ξ| − 1)2 sin |ξ| − |ξ|
+ + (sin |ξ| − |ξ|)|ξ|
−c2 + c2 |ξ| + c21
2
|ξ|2 |ξ|
a2 = = .
(1 − c2 |ξ|2 )2 + c21 |ξ|2 (sin |ξ|)2 + (1 − cos |ξ|)2
Bibliography
[AKW93] M. Austin, P. S. Krishnaprasad, and L.-S. Wang: Almost Poisson integration of rigid
body systems. J. of Comp. Phys., 107:105–117, (1993).
[AM78] R. Abraham and J. E. Marsden: Foundations of Mechanics. Addison-Wesley, Reading,
MA, Second edition, (1978).
[AN90] A. I. Arnold and S.P. Novikov: Dynomical System IV. Springer Verlag, Berlin Heidel-
berg, (1990).
[Arn89] V. I. Arnold: Mathematical Methods of Classical Mechanics. Springer-Verlag, GTM
60, Berlin Heidelberg, Second edition, (1989).
[CFSZ08] E. Celledoni, F. Fassò, N. Säfström, and A. Zanna: The exact computation of the
free rigid body motion and its use in splitting methods. SIAM J. Sci. Comput., 30(4):2084–
2112, (2008).
[CG93] P. E. Crouch and R. Grossman: Numerical integration of ordinary differential equa-
tions on manifolds. J. Nonlinear. Sci., 3:1–33, (1993).
[CS90] P. J. Channell and C. Scovel: Symplectic integration of Hamiltonian systems. Nonlin-
earity, 3:231–259, (1990).
[CS91] P. J. Channel and J. S. Scovel: Integrators for Lie–Poisson dynamical systems. Physica
D, 50:80–88, (1991).
[Fen85] K. Feng: On difference schemes and symplectic geometry. In K. Feng, editor, Pro-
ceedings of the 1984 Beijing Symposium on Differential Geometry and Differential Equa-
tions, pages 42–58. Science Press, Beijing, (1985).
[Fen86] K. Feng: Symplectic geometry and numerical methods in fluid dynamics. In F. G.
Zhuang and Y. L. Zhu, editors, Tenth International Conference on Numerical Methods in
Fluid Dynamics, Lecture Notes in Physics, pages 1–7. Springer, Berlin, (1986).
[FQ87] K. Feng and M.Z. Qin: The symplectic methods for the computation of Hamiltonian
equations. In Y. L. Zhu and B. Y. Guo, editors, Numerical Methods for Partial Differential
Equations, Lecture Notes in Mathematics 1297, pages 1–37. Springer, Berlin, (1987).
[FQ91] K. Feng and M.Z. Qin: Hamiltonian algorithms for Hamiltonian systems and a com-
parative numerical study. Comput. Phys. Comm., 65:173–187, (1991).
[FWQ90] K. Feng, H.M. Wu, and M.Z. Qin: Symplectic difference schemes for linear Hamil-
tonian canonical systems. J. Comput. Math., 8(4):371–380, (1990).
[FWQW89] K. Feng, H. M. Wu, M.Z. Qin, and D.L. Wang: Construction of canonical dif-
ference schemes for Hamiltonian formalism via generating functions. J. Comput. Math.,
7:71–96, (1989).
[Ge91] Z. Ge: Equivariant symplectic difference schemes and generating functions. Physica
D, 49:376–386, (1991).
[GM88] Z. Ge and J. E. Marsden: Lie–Poisson–Hamilton–Jacobi theory and Lie–Poisson in-
tegrators. Physics Letters A, pages 134–139, (1988).
[HV06] E. Hairer and G. Vilmart: Preprocessed discrete Moser–Veselov algorithm for the full
dynamics of the rigid body. J. Phys. A, 39:13225–13235, (2006).
[Kar04] B. Karasözen: Poisson integrator. Math. Comput. Modelling, 40:1225–1244, (2004).
[Lie88] S. Lie: Zur theorie der transformationsgruppen. Christiania, Gesammelte Abh., Christ.
Forh. Aar., 13, (1988).
548 Bibliography
[LQ95a] S. T. Li and M. Qin: Lie–Poisson integration for rigid body dynamics. Computers
Math. Applic., 30:105–118, (1995).
[LQ95b] S. T. Li and M. Qin: A note for Lie–Poisson– Hamilton–Jacobi equation and Lie–
Poisson integrator. Computers Math. Applic., 30:67–74, (1995).
[McL93] R.I. McLachlan: Explicit Lie–Poisson integration and the Euler equations. Physical
Review Letters, 71:3043–3046, (1993).
[MR99] J. E. Marsden and T. S. Ratiu: Introduction to Mechanics and Symmetry. Number 17
in Texts in Applied Mathematics. Springer-Verlag, Berlin, Second edition, (1999).
[MRW90] J.E. Marsden, T. Radiu, and A. Weistein: Reduction and hamiltonian structure on
dual of semidirect product Lie algebra. Contemporary Mathematics, 28:55–100, (1990).
[MS95] R. I. McLachlan and C. Scovel: Equivariant constrained symplectic integration. J.
Nonlinear. Sci., 5:233–256, (1995).
[MS96] R. I. McLachlan and C. Scovel: A Survey of Open Problems in Symplectic Integration.
In J. E. Mardsen, G. W. Patrick, and W. F. Shadwick, editors, Integration Algorithms and
Classical Mechanics, pages 151–180. American Mathematical Society, New York, (1996).
[MV91] J. Moser and A. P. Veselov: Discrete versions of some classical integrable systems and
factorization of matrix polynomials. Communications in Mathematical Physics, 139:217–
243, (1991).
[MW83] J.E. Marsden and A. Weinstein: Coadjoint orbits, vortices and Clebsch variables for
incompressible fluids. Phys D, 7: (1983).
[MZ05] R.I. McLachlan and A. Zanna: The discrete Moser–Veselov algorithm for the free
rigid body. Foundations of Computational Mathematics, 5(1):87–123, (2005).
[Olv93] P. J. Olver: Applications of Lie Groups to Differential Equations. GTM 107. Springer-
Verlag, Berlin, Second edition, (1993).
[Qin89] M. Z. Qin: Cononical difference scheme for the Hamiltonian equation. Mathematical
Methodsand in the Applied Sciences, 11:543–557, (1989).
[TF85] H. Tal-Fzer: Spectral method in time for hyperbolic equations. SIAM J. Numer. Anal.,
23(1):11–26, (1985).
[Ves88] A.P. Veselov: Integrable discrete-time systems and difference operators. Funkts. Anal.
Prilozhen, 22:1–33, (1988).
[Ves91] A.P. Veselov: Integrable maps. Russian Math. Surveys,, 46:1–51, (1991).
[Wan91] D. L. Wang: Symplectic difference schemes for Hamiltonian systems on Poisson
manifolds. J. Comput. Math., 9(2):115–124, (1991).
[ZQ94] W. Zhu and M. Qin: Poisson schemes for Hamiltonian systems on Poisson manifolds.
Computers Math. Applic., 27:7–16, (1994).
[ZS07a] R.van Zon and J. Schofield: Numerical implementation of the exact dynamics of free
rigid bodies. J. of Comp. Phys., 221(1):145–164, (2007).
[ZS07b] R.van Zon and J. Schofield: Symplectic algorithms for simulations of rigid body
systems using the exact solution of free motion. Physical Review E, 50:5607, (2007).
Chapter 13.
KAM Theorem of Symplectic Algorithms
Among the various kinds of equations of mathematical physics, only a few can be
integrated exactly by quadrature and the rest are unsolvable. However, even an ap-
proximate solution is also valuable in many scientific and engineering problems. In a
wide range of applications, the most powerful and perhaps the only practically feasible
approximation is the numerical method — this is the case, especially in the computer
era. A question arises accordingly: Whether a numerical method can reflect the real
information of exact solutions of original problems properly or simulate accurately?
To a problem described by time evolutionary equations, the solutions can often be
represented by a flow (or semi-flow), which is locally defined on a phase space. Curves
on the phase space which are invariant under the action of the flow (or semi-flow) are
called invariant curves (or positively invariant curves) of the flow (or semi-flow). There
is a natural correspondence between the solutions of the equations and the invariant
curves (or positively invariant curves) of the flow (or semi-flow). The invariant curves,
or positively invariant curves, are called solution curves of the equations. The quali-
tative analysis concerns with problems about understanding topological structures of
the solution curves and their limit sets, which are often sub-manifolds of the phase
space. The aim of the numerical method, in principle, not only pursues an optimal
quantitative approximation to the real solution of the considered problem locally but
also preserves as well as possible the topological and even geometrical properties of
550 13. KAM Theorem of Symplectic Algorithms
the solution curves and their limit sets globally. The latter constitutes the main content
of qualitative analysis of the numerical method.
Qualitative analysis becomes important in the study of numerical methods be-
cause instability phenomena take place very often even in the numerical simulations
of very stable systems. Numerical treatments of stiff problems show that explicit
methods have a severe time-step restriction, which lead to Dahlquist’s pioneering
work about A-stability[Dah63] . Various notions of stability for numerical methods have
been established since then, classifying different types of stable methods for differ-
ent problems. The celebrated linear stability theory (A-stability, A(α)-stability, and
L-stability)[Wid76,Ehl69] is based on the scalar linear equation1
ẏ = λy (1.1)
and turns out to be powerful for the numerical study of all linear-dissipation-dominated
problems. G-stability, which was also developed by Dahlquist[Dah75] , is characterized
by retaining the contractivity property of any two solutions of nonlinear “contractive”
systems
ẏ = f (y). (1.2)
Here, the “contractivity” of the system (1.2) is defined by the condition
topological structures of most dynamic trajectories of typical stable systems (e.g., dis-
sipative systems having motion stability and Morse–Smale systems and Axiom A sys-
tems having structure-stability)[SH96,Li99] . However, this remarkable advantage of the
conventional methods does not carry over to conservative systems. Most of the sta-
ble methods introduce artificial dissipation into conservative systems. They produce
illusive attractors and therefore destroy the qualitative character of the conservative
systems even if sufficiently small step sizes of the numerical method are used. The
approximate solutions of conservative systems ask for new numerical methods which
require more stringent stability.
Geometric numerical integration theory for conservative systems has been devel-
oped rapidly in recent twenty years. The monographs[SSC94,HLW02,FQ03,LR05] summarize
the main developments and important results of this theory. Qualitative behavior of ge-
ometric integrators has been investigated by many authors[Sha99,Sha00b,HL97,Sto98a,HLW02] .
For symplectic integrators applied to Hamiltonian systems, some stability results, ei-
ther in the spirits of the KAM theory or based on the backward analysis, have been
well established[Sha99,Sha00b,HL97,CFM06,DF07] . The typical stable dynamics of Hamilto-
nian systems, e.g., quasi-periodic motions and their limit sets — minimal invariant
tori, can be topologically preserved and quantitatively approximated by symplectic
discretizations. In this chapter, we give a review about these results. For more details,
readers refer to the relevant references[Sha99,Sha00b,HLW02,CFM06,DF07] .
where ∂ denotes the gradient operator with respect to p. Under the mapping S0 ,
the phase space I × Tn is completely foliated into invariant n-tori {p} × Tn ,
p ∈ I. On each torus, the iterations of S0 are linear with frequencies ω = ω(p).
This is a typical integrable case. When a perturbation h(p, q) is added to H0 , i.e.,
H(p, q) = H0 (p) + h(p, q), (2.1) does not define an integrable mapping generally.
However, KAM theorem shows that the perturbed mapping S still exhibits to a large
extent the integrable behavior in the phase space if the frequency map ω is nondegen-
erate in some sense (see[Arn63,AA89,Arn89,Kol54b,Mos62] for Kolmogorov’s nondegeneracy
and[CS94,Rüs90] for weak nondegeneracy) and the perturbation h is sufficiently small in
some function space. In this chapter, we consider the following nondegeneracy condi-
tion for ω : I → Ω:
for some 0 < θ ≤ Θ. Here I and Ω are the domains of action variables and the
corresponding frequency values respectively. We always assume that I and Ω are open
in Rn and ω is analytic and can be analytically extended to some complex domain,
say I + r, of the real domain I, where r is the extension radius. We assume (2.3) is
satisfied for p1 , p2 ∈ I + r with |p1 − p2 | ≤ r. Note that this nondegeneracy condition
implies that the frequency map ω is invertible in any ball of radius r and centered in I,
which is stronger than the standard Kolmogorov’s nondegeneracy assumption of the
following (this was already noticed by Pöschel in[Pös82] ,
n
n
with some constants γ > 0, τ > 0, where k, ω = kj ωj and |k| = |kj | for
j=1 j=1
integers k ∈ Zn .
We introduce some notations. For an open or closed set I ⊂ Rn and for a ≥ 0,
denote by C a (I ×Tn ), the class of isotropic differentiable functions of order a defined
13.2 Mapping Version of the KAM Theorem 553
for u ∈ C ν1 ,ν2 (I × Tn ), where σρ denotes the partial stretching (x, y) → (ρx, y) for
(x, y) ∈ I × Tn . Note that the following relation between these two norms is valid
for 0 < ρ ≤ 1:
uν1 ,ν2 ;ρ ≤ uν1 ,ν2 ≤ ρ−ν1 uν1 ,ν2 ;ρ , (2.7)
where we dropped the domains to simplify the notations.
Take Ω = ω(I) and denote by Ωγ the set of those frequencies, in Ω, which satisfy
the diophantine condition (2.5) for given γ > 0 and whose distance to the boundary
#
of Ω is at least equal to 2γ. The set Ωγ is a Cantor set2 and the difference Ω \ Ωγ
γ>0
is a zero set if τ > n + 1. Therefore Ωγ is large for small γ.
The main results of this section are stated as follows:
Theorem 2.1. Given positive integer n and real number τ > n+1, consider mapping
S defined in phase space I ×Tn by (2.1) with H( p, q) = H0 (
p)+h( p, q), where H0 is
p, q) belongs to the Whitney’s class C αλ+λ+τ (I ×
analytic in I + r with r > 0 and h(
Tn ) for some λ > τ + 1 and α > 1,
i
α∈/Λ= + j : i, j ≥ 0 integer .
λ
Suppose the frequency map ω = ∂H0 : I → Ω satisfies the nondegeneracy condition
(2.3) for p1 , p2 ∈ I + r with |p1 − p2 | ≤ r where the constants θ and Θ satisfy
0 < θ ≤ Θ, then there exists a positive
constant
δ0 , depending only on n, τ , λ and α,
1
such that for any 0 < γ ≤ min 1, rΘ , if
2
S ◦ Φ = Φ ◦ R, (2.9)
2◦ If Ω is a bounded open set of type D in the Arnold’s sense3 , then we have the
following measure estimate
−n
mEγ ≥ 1 − c4 θΘ −1 γ m E, (2.10)
Theorem 2.2. If the frequency map ω satisfies the nondegeneracy condition (2.4),
then the conclusions of Theorem 2.1 are still true with the same estimates (2.10) –
(2.12) under the following smallness condition for h,
Remark 2.3. The above two theorems are stated for the case when h is finitely many
times differentiable. If h is infinitely many times differentiable or analytic, we have the
following conclusions, which are easily derived by similar remarks to those of[Pös82] .
1◦ If h ∈ C ∞ (I × Tn ), then ωγ ∈ C ∞ (Iγ ) and Φ ∈ C ∞ (Iγ × Tn ) and the
estimates (2.8) hold for any β ≥ α.
2◦ If h ∈ C ω (I × Tn ), then we have further Φ ∈ C ∞,ω (Iγ , Tn ) under an
additional smallness condition for δ0 which also depends on the radius of analyticity of
h with respect to angle variables. Here C ω denotes the class of real analytic functions.
and
Iρ = ρ−1 I = {x ∈ Rn |ρx ∈ I}. (2.17)
For the time being, ρ is regarded as a free parameter. F0 (x) is real analytic in
Iρ + rρ and f belongs to the class C a (Iρ × Tn ) where rρ = ρ−1 r, a = αλ + λ + τ .
So the new mapping T satisfies the assumptions of Theorem 2.1 in which only I,
r, H, H0 , and h are replaced by Iρ , rρ , F , F0 , and f respectively. Accordingly, the
frequency map of the integrable mapping associated to the generating function F0
turns into
" (x) = ∂F0 (x), x ∈ Iρ ,
ω
and the nondegeneracy condition for the mapping turns out to be
ρθ |x1 − x2 | ≤ |"
ω (x1 ) − ω
" (x2 )| ≤ ρΘ |x1 − x2 | (2.18)
1
From now on, we fix ρ = γΘ−1 . Then the assumption 0 < γ ≤ rΘ in Theorem
2
1
2.1 implies that 0 < ρ ≤ r. Hence, rρ ≥ 2. Let Iρ∗ be the set of points in Iρ with the
2
distance to its boundary at least one and let
" −1 (Ωγ ) ∩ Iρ .
Iρ;γ = ω (2.19)
and
γμ |x1 − x2 | ≤ |"
ω (x1 ) − ω
" (x2 )| ≤ γ |x1 − x2 | , μ = θΘ−1 (2.21)
Uj = Iρ × Tn + (4sj , 4sj )
where cb is a positive constant only depending on b, n and s0 but not depending on the
domain Iρ and hence not depending on the parameter ρ. Moreover, we may require fj
to be 2π-periodic in the last n variables. In (2.23), | · |Uj denotes the maximum norm
of analytic functions on the complex domain Uj .
c. We give the KAM iteration process which essentially follows Pöschel [Pös82] in
the Hamiltonian system case. For each fj , we define a mapping Tj : (x, y) → ( x, y)
by
= x − ∂2 Fj (
x x, y), y = y + ∂1 Fj ( x, y) (2.24)
with Fj (
x, y) = F0 (
x)+fj (x, y). For each j, the function Fj (
x, y) is well-defined and
real analytic on Uj if 4sj ≤ rρ = ρ−1 r — this inequality is satisfied for j = 0, 1, · · ·
1
if we choose 0 < s0 ≤ 4−1 noting that 0 < γ < rΘ . We can show that each Tj
2
for j ≥ 0 is a well-defined analytic mapping on a domain of complex extension of
the phase space Iρ × Tn , which is appropriate for the KAM iterations if h is bounded
by (2.8) with a sufficiently small δ0 > 0. It follows from (2.23) that Tj converges to
T in C b−1−κ -norm for any κ > 0 on some subdomain Iρ∗ × Tn of the phase space
Iρ × Tn , where T is well-defined. The central problem is to find transformations
Φj and integrable rotations Rj , defined on a sequence of nested complex domains
that intersect a nonempty Cantor set, say I"ρ;γ × Tn , such that the following holds as
j → ∞ on I"ρ;γ × Tn in some Whitney’s classes,
" and R
where Φ " are well-defined on I"ρ;γ × Tn . In this case, we have
"=Φ
T ◦Φ " ◦R
" on I"ρ;γ × Tn . (2.26)
S◦Φ=Φ◦R on Iγ × Tn ,
where
Iγ = ρI"ρ;γ = {p ∈ Rn | ρ−1 p ∈ I"ρ;γ }
13.2 Mapping Version of the KAM Theorem 557
is a Cantor set of I. In fact, due to the nondegeneracy of the frequency map ω in the
sense of (2.3), we may keep the frequencies prescribed by (2.5) fixed in the above
approximation process. As a result, we have ωγ (Iγ ) = Ωγ , where ωγ is the frequency
map of the integrable rotation R on Iγ ×Tn . This is just the conclusion (1) of Theorem
2.1.
The construction of Φj and Rj uses the KAM iteration, which is described as
follows.
Assume
|fj − fj−1 | ∼ εj , j = 1, 2, · · · , (2.27)
where εj is a decreasing sequence of positive numbers . Suppose we have already
found a transformation Φj and a rotation Rj with frequency map ω (j) such that
Cj = Rj−1 ◦ Φ−1
j ◦ Tj ◦ Φj (2.28)
satisfies
|Cj − I| ∼ εj+1 . (2.29)
Then, we construct a transformation Ψj and a new rotation Rj+1 with frequency map
ω (j+1) such that
Φj+1 = Φj ◦ Ψj (2.30)
and (2.29) is also true for the next index j + 1 with Cj+1 defined by (2.28) in which
j is replaced by j + 1. As was remarked by Pöschel in[Pös82] , “for this procedure
to be successful it is essential to have precise control over the various domains of
definition”.
We define transformation Ψj : (ξ, η) → (x, y) implicitly with the help of a gener-
ating function ψj by
Bj = Rj−1 ◦ Φ−1
j ◦ Tj+1 ◦ Φj . (2.32)
Bj is near identity and is assumed to be given implicitly from its generating function
bj by
= x − ∂2 bj (
x x, y), y = y + ∂1 bj (
x, y). (2.33)
The function ψj is then determined from bj by the following homological equation
where "bj (x, y) = bj (x, y) − [bj ](x) with [bj ] being the mean value of bj with respect
to the angle variables over Tn . Define
= x,
x y = y + ω (j+1) (x). (2.36)
Ψ−1
j ◦ Rj ◦ Bj ◦ Ψj = Rj+1 ◦ Cj+1 .
Formal calculations similar to those in[Pös82] show that (2.29) is valid if we replace j
by j + 1.
We do not solve the Equation (2.34) exactly. Instead, we will solve an approximate
equation by truncating the fourier expansion of "bj with respect to angle variables to
some finite order so that “only finitely many resonances remain, and we obtain a real
analytic solution ψj defined on an open set”[Pös82] . This idea was first successfully
used by Arnold[Arn63,AA89] .
For an earnest proof, we need more precise arguments by carefully controlling the
domains of definition of functions and mappings in the iterative process. Readers can
refer to[Sha00a] for details.
where H( p, q) = H0 (p) + h(p, q) and ω( p). Under the assumptions of
p) = ∂H0 (
Theorem 2.1 (Theorem 2.2) for H, if h satisfies the smallness condition of Theorems
2.1 (Theorem 2.2), then the corresponding conclusions of the theorem are still valid
for St (0 < t ≤ 1), only with the following remarks:
1◦ Ωγ is replaced by
tγ
Ωt,γ = ω ∈ Ω∗ : eik,tω − 1 ≥ for k ∈ Zn \ {0} , (2.38)
|k|τ
where Ω∗ denotes the set of points in Ω with distance to its boundary at least equal to
2γ. Accordingly, Iγ , ωγ , Φ, and R are replaced by It,γ , ωt,γ , Φt and Rt ,respectively.
2◦ If Ω is a bounded open set of type D in the Arnold’s sense[Arn63] , then we have
the following Lebesgue measure estimate
for t ∈ (0, 1], with constant D only depending on n, τ and the geometry of Ω. So in
this case, Ωt,γ is still a large Cantor set in Ω when γ is small enough.
3◦ If h ∈ C ∞ (B × Tn ), then ωγ,t ∈ C ∞ (Bγ,t ) and Φt ∈ C ∞ (Bγ,t × Tn )
which satisfy the estimates (2.11) for any β ≥ α.
4◦ If h is analytic with the domain of analyticity containing
( )
S(r, ρ) = (p, q) ∈ C2n : |p − p | < r, |Im q| < ρ with p ∈ B and Re q ∈ Tn
for some r > 0 and ρ > 0 (Re q and Im q denote the real and imaginary parts of q
respectively) and if h satisfies
hr,ρ = sup |h(p, q)| ≤ δ0 γ 2 θ2 Θ−3 (2.40)
(p,q)∈S(r,ρ)
for some sufficiently small δ0 > 0 depending on n, τ , r and ρ, then all the con-
clusions of Theorem 2.1 (Theorem 2.2) are still true with ωγ,t ∈ C ∞ (Bγ,t ), Φt ∈
C ∞,ω (Bγ,t × Tn ) and the estimate (2.11) holds for any β ≥ 0.
We have presented the results about the existence of differentiable foliation struc-
tures in the sense of Whitney of invariant tori for nearly integrable symplectic map-
pings and for mappings with small twists. Such a result was proved first by Lazutkin
in 1974[Laz74] for planar twist maps and was generalized to higher dimensions by
Svanidze in 1980[Sva81] . For the case of Hamiltonian flows of arbitrary dimensions,
the generalizations were given by J. Pöschel in 1982[Pös82] , Chierchia and Gallavotti
in 1982[CG82] . The perturbation and measure estimates in terms of γ were studied
by Rüssmann in 1981[Rüs81] , Svadnidze in 1981[Sva81] , Neishtadt in 1982[Nei82] and
Pöschel in 1982[Pös82] . The estimates in terms of θ and Θ were given by Shang in
2000[Sha00a] , which are also crucial in the small twist mapping case.
where D is a connected bounded, open subset of R2n ; x and y are both n-dimensional
Euclidean coordinates with ẋ and ẏ the derivatives of x and y with respect to the time
“t” respectively; K : D → R1 is the Hamiltonian.
A symplectic algorithm that is compatible with the system (3.1) is a discretization
scheme such that, when applied to the system (3.1), it uniquely determines one param-
eter family of symplectic step-transition maps GtK that approximates the phase flow
t
gK in the sense that
1
lim s GtK (z) − gK t
(z) = 0 for any z = (x, y) ∈ D (3.2)
t→0 t
for some s ≥ 1, here t > 0 is the time-step size of the algorithm and s, the largest
integer such that (3.2) holds, is the order of accuracy of the algorithm approximating
the continuous systems. Note that the domain in which GtK is well-defined, say D " t,
depends on t generally and converges to D as t → 0 — this means that any z ∈ D
may be contained in D " t when t is sufficiently close to zero.
From (3.2), we may assume
GtK (z) = gK
t
(z) + ts RK
t
(z), (3.3)
where
1 t
t
RK (z) = s
GK (z) − gK
t
(z)
t
is well-defined for z ∈ D" t ⊂ D and has the limit zero as t → 0 for z ∈ D. Below, we
prove the main results of this chapter by simply regarding the approximation GtK to
t
the phase flow gK of the above form as a symplectic discretization scheme of order s.
We assume that the system (3.1) is integrable. That is, there exists a system of
action-angle coordinates (p, q) in which the domain D can be expressed as the form
B × Tn and the Hamiltonian depends only on the action variables, where B is a
connected bounded, open subset of Rn and Tn the standard n-dimensional torus. Let
us denote by Ψ : B × Tn → D the coordinate transformation from (p, q) to (x, y),
then Ψ is a symplectic diffeomorphism from B × Tn onto D. The new Hamiltonian
only depends on p. Therefore, in the action-angle coordinates (p, q), (3.1) takes the
simple form
∂H
ṗ = 0, q̇ = ω(p) = (p) (3.5)
∂p
and the phase flow gHt
is just the one parameter group of rotations (p, q) → (p, q +
tω(p)) which leaves every torus {p} × Tn invariant.
13.3 KAM Theorem of Symplectic Algorithms for Hamiltonian Systems 561
Assume K is analytic and, without loss of generality, assume the domain of ana-
lyticity of K contains the following open subset of C2n
( )
Dα0 = z = (x, y) ∈ C2n : d(z, D) < α0 , (3.6)
denotes the distance from the point z ∈ C2n to the set D ⊂ C2n in which |z| =
max |zj | for z = (z1 , · · · , z2n ). Also, we assume that Ψ extends analytically to the
1≤j≤2n
following complex domain
( )
S(r0 , ρ0 ) = (p, q) ∈ C2n : d(p, B) < r0 , Re q ∈ Tn , |Im q| < ρ0 (3.7)
Ψ−1 ◦ gK
t
◦ Ψ = gH
t
(3.8)
ξ
t ∈ [0, δ1 ], where V1 + denotes the union of all complex open balls centered in V1
2
ξ
with radius . Since M (z, t) is continuous and positive for (z, t) ∈ V1 × [0, δ1 ], there
2
exists a constant M 0 > 0 which is an upper bound of M (z, t) on V1 × [0, δ1 ]. Let
D
ξ
δ2 = min 1, δ1 , . Then for t ∈ [0, δ2 ], GtK maps V1 into Dα0 and hence
4M0
G" t = Ψ−1 ◦ Gt ◦ Ψ is well-defined on U1 . The real analyticity of the map follows
K K
from the real analyticity of Ψ and K. To verify Equation (3.9) , we first note that the
3ξ
analyticity of Ψ−1 on V1 + ⊂ Dα0 implies that
4
−1
∂Ψ
(z) ≤ M1
∂z
3ξ
for all z ∈ V1 + with some constant M1 > 0,and then Taylor formula gives
4
Ψ(p, q) ∈ V1 and
t ξ
RK (Ψ(p, q)) = GtK (Ψ(p, q)) − gK
t
(Ψ(p, q)) ≤ M0 ts+1 ≤
4
for (p, q) ∈ U1 and t ∈ [0, δ2 ]. Therefore,
" t (p, q) − g t (p, q)| = |Ψ−1 (g t (Ψ(p, q)) + Rt (Ψ(p, q))) − Ψ−1 (g t (Ψ(p, q)))|
|G K H K K K
≤ 2nM1 M0 ts+1 .
The above lemma shows that G " t is an approximant to the one parameter group
K
of integrable rotations gH up to order ts+1 as t approaches zero. To apply Theorem
t
2.4, we need to verify the exact symplecticity of G" t so that it can be expressed by
K
globally defined generating function. Because Ψ is not necessarily exact symplectic,
" t = Ψ−1 ◦ Gt ◦ Ψ is not trivially observed.
the exact symplecticity of G K K
Lemma 3.2. Let G be an exact symplectic mapping of class C 1 from D into R2n
where D is an open subset of R2n and let Ψ be a symplectic diffeomorphism from
B × Tn onto D. Then Ψ−1 ◦ G ◦ Ψ is an exact symplectic mapping in the domain in
which it is well-defined.
p, q) = Ψ−1 ◦ G ◦ Ψ(p, q) and let γ be any given closed curve in the
Proof. Let (
domain of definition of G" =: Ψ−1 ◦ G ◦ Ψ, which is an open subset of B × Tn . The
"
exact symplecticity of G will be implied by[Arn89]
- -
I(γ) = p d q − p d q = 0. (3.10)
γ γ
x, y) = Ψ(
Now we verify (3.10). Let (x, y) = Ψ(p, q) and ( p, q). Then (
x, y) =
d y − γ x d y = 0, where x,
G(x, y). Since G is an exact symplectic, we have γ x
13.3 KAM Theorem of Symplectic Algorithms for Hamiltonian Systems 563
, y are considered as functions of (p, q), which vary over γ. Therefore, with these
y, x
conventions and with γ = Ψ−1 ◦ G ◦ Ψ(γ),
- - - -
I(γ) = p d q − x d y + x d y − p d q
γ γ γ γ
- - - -
= pdq − xdy + xdy − pdq
γ Ψ(γ ) Ψ(γ) γ
- -
= pdq − x d y. (3.11)
γ −γ Ψ(γ )−Ψ(γ)
Note that G is exact and hence it is homotopic to the identity. This implies that Ψ−1 ◦
G ◦ Ψ is homotopic to the identity too. So γ and γ belong to the same homological
class in the fundamental group of the manifold B × Tn . Therefore, one may find a
2-dimensional surface, say σ, in the phase space B × Tn , which is bounded by γ and
γ. Ψ(σ) is then a 2-dimensional surface in D bounded by Ψ(γ ) and Ψ(γ). By stokes
formula and from (3.11), we get
- -
I(γ) = dp ∧ dq − d x ∧ d y,
σ Ψ(σ)
r ρ
( p, q) in S 0 , 0 and then taking the maximal norm of the integration
p0 , q0 ) to (
4 4
r0 ρ0
for (p, q) over S , , we obtain the estimate
4 4
Proof. Now the analytic version of Theorem 2.4 can be applied to St = G " t . The
K
conditions required by Theorem 2.4 are satisfied clearly according to the assumptions
of Theorem 2.1. For example, the nondegeneracy of the integrable system in the sense
of Kolmogorov means that the frequency map ω : B → Rn is nondegenerate and
therefore, there exists positive constants θ ≤ Θ such that ω satisfies (2.4) with some
positive numbers r ≤ r0 . We assume r = r0 here without loss of generality. In
Theorem 2.4, the function h is replaced by ts ht which satisfies the estimate (2.40)
r ρ
with r = 0 and ρ = 0 if we choose
4 4
D
s 2nM L −1 32
γ = γt =: Γ td , with 0 < d ≤ and Γ = θ Θ (3.14)
2 δ0
and if t is sufficiently small, where δ0 is the bound given by (2.40) of Theorem 2.4.
1
It is clear that the so chosen γ satisfies the condition γ ≤ min 1, rΘ required by
2
Theorem 2.4 for t sufficiently close to zero. By Theorem 2.4, we then have the Cantor
sets Bt = Bγ,t ⊂ B and Ωt = Ωγ,t ⊂ ω(B), a surjective map ωt = ωγ,t : Bt → Ωt
of class C ∞ and a symplectic mapping Φt : Bt × Tn → Rn × Tn of class C ∞,ω ,
in the sense of Whitney, such that the conclusions (1) – (4) of Theorem 2.4 hold with
" t fill out a set Et = Eγ,t = Φt (Bt × Tn ) in
γ = Γ td . From (2.10), invariant tori of G K
phase space E = B × T with measure estimate
n
−n d
mEt ≥ 1 − c4 Γ θΘ −1 t mE. (3.15)
From (2.11), with the notice of (2.7) and the fact that
In the last inequality of (3.16), we have used the estimate (3.13) for ht . From (2.11),
we also get
with constant "c4 > 0 not depending on γ and t, where Dγ,t = Ψ(Eγ,t ) with Eγ,t =
Φt (Bγ,t × Tn ) and with Bγ,t being the subset of B as indicated above. Note that
Bγ,t is a closed subset of Bt and Dγ,t a closed subset of Dt if t is sufficiently small.
Moreover, the estimate
c8 γ −(2+β) Θ1+β · ts
Ψt − Ψβ,βλ;Bγ,t ×Tn ≤ " (3.20)
and
c8 γ −(2+β) Θ2+β · ts
ωt − ωβ+1;Bγ,t ≤ " (3.21)
hold for any β ≥ 0 with " c8 > 0 not depending on γ and t. The conclusions (1)
−n
and (2) of the last part of Theorem 3.4 are proved if we set c1 = " c4 θΘ −1 and
c2 = "
c8 ·max(Θ1+β , Θ2+β ). From (3.19), it follows that for a sufficiently small γ > 0,
Dγ,t has a positive Lebesque measure. From (2.12), it follows that for any ω ∗ ∈ Ωγ,t ,
there exists p∗ ∈ B and P ∗ ∈ Bγ,t such that ω(p∗ ) = ωt (P ∗ ) = ω ∗ and
−1 s
|P ∗ − p∗ | ≤ 2nM Lc6 c7 γθΘ−1 ·t ,
Corollary 3.5. Under the assumptions of the above theorem, there exists n functions
F1t , · · · , Fnt which are defined on the Cantor set Dt and of class C ∞ in the sense of
Whitney such that:
1◦ F1t , · · · , Fnt are functionally independent and in involution (i.e., the Poisson
bracket of any two functions vanishes on Dt );
2◦ Every Fjt (j = 1, · · · , n), is invariant under the difference scheme and the
invariant tori are just the intersection of the level sets of these functions;
3◦ Fjt (j = 1, · · · , n) approximate n independent integrals in involution of the
integrable system, with a suitable order of accuracy with respect to the time-step t
which will be explained in the proof.
where Rt is the integrable rotation (p, q) → (p, q + tωt (p)) and admits n invariant
functions, say, p1 , · · · , pn , analytically defined on Bt × Tn . Let
Fit = pi ◦ Ψ−1
t , i = 1, · · · , n,
then they are well-defined on the Cantor set Dt and of class C ∞ in the sense of
Whitney due to the C ∞ -smoothness of Ψ−1 t on Dt . Moreover, we easily verify by
(3.22) that
Fit ◦ GtK = Fit , i = 1, · · · , n,
and this means that Fit (i = 1, · · · , n) are n invariant functions of GtK . These n invari-
ant functions are functionally independent because pi (i = 1, · · · , n) are functionally
independent and Ψt is a diffeomorphism. The claim that Fit and Fjt are in involution
for 1 ≤ i, j ≤ n simply follows from the fact that pi and pj are in involution and Ψt
is symplectic. Note that the Poisson bracket is invariant under symplectic coordinate
transformations. Finally, it is observed from the proof of Theorem 3.4 that for each of
j = 1, · · · , n, Fjt approximates
Fj = pj ◦ Ψ−1
s
as t → 0, with the order of accuracy equal to ts−(2+β)d 0 < d < is given on
2+β
the set Dt (note that this set depends also on d by definition) and equal to ts on Dγ,t ,
a subset of Dt , in the norm of the class C β for any given β ≥ 0. It is clear that the
functions Fj (j = 1, · · · , n) are integrals of the integrable system and that any two of
them are in involution by the symplecticity of Ψ−1 . Corollary 3.5 is then proved.
568 13. KAM Theorem of Symplectic Algorithms
with some γ > 0 and τ > 0, where u, v denotes the inner product of vectors u and
v in Rn . Note that t > 0 may be arbitrarily small.
For any fixed ω ∈ Rn , even if it is a diophantine vector, there exists some t in any
small neighborhood of the origin such that (4.1) does not hold for any γ > 0 and any
τ > 0. In fact, one can choose t to satisfy the resonance relation
for some 0 = k ∈ Zn . In the next section, we will show that such t forms a dense set
in R.
We note that a one-step algorithm, when applied to system of differential equa-
tions, can be regarded as a perturbation of the phase flow of the system. On the other
hand, according to Poincaré, arbitrarily small perturbations in the generic case may
destroy those resonant invariant tori of an integrable system. Therefore, to simulate
the invariant torus with a given frequency of some Hamiltonian system by symplectic
algorithms, one is forced to be very careful to select step sizes, say, to keep them away
from some dense set.
Some questions arise: is it possible to simulate an invariant torus of an integrable
system by symplectic algorithms? If possible, how does one select the step sizes and
what structure does the set of those admitted step sizes have? In this paper, we try to
answer these questions.
Lemma 4.1. For any ω ∈ Rn , there exists a dense subset, say D(ω), of R such that
for any t ∈ D(ω), the resonance relation (4.2) holds for some 0 = k ∈ Zn .
which is clearly dense in R and the resonance relation (4.2) holds for any t ∈ D(ω).
The proof of the lemma is completed.
13.4 Resonant and Diophantine Step Sizes 569
Definition 4.2. D(ω) is called the resonant set of step sizes with respect to the fre-
quency ω ∈ Rn . Any t ∈ D(ω) is called a resonant step size with respect to ω.
We denote by Iλ,μ,τ (ω), the set of numbers τ satisfying (4.4) for given constants
λ > 0, μ and τ > 0. Then, Iλ,μ,τ (ω) is a subset of R which is far away from resonance
with respect to ω. For this set, we have:
Lemma 4.4. For any nonresonant frequency ω ∈ Rn , and for any λ > 0, any μ and
any τ > 0, the set Iλ,μ,τ (ω) is nowhere dense and closed in R. Moreover, if μ > 1
and τ > n, then we have
meas R \ Iλ,μ,τ (ω) ≤ cλ, (4.5)
Proof. The nowhere denseness and the closedness of Iλ,μ,τ (ω) follow from the fact
that the complement of the set is both open and dense in R for any λ > 0, μ and
τ > 0. It remains to prove (4.5). Since
570 13. KAM Theorem of Symplectic Algorithms
, ' /
2πl λ
R \ Iλ,μ,τ (ω) = t ∈ R : t − < ,
k, ω lμ |k|τ
0<l∈Z
0=k∈Zn
we have
∞
1
2λ 1
meas R \ Iλ,μ,τ (ω) ≤ ≤ 2λ · .
lμ |k|τ lμ |k|τ
0<l∈Z l=1 0=k∈Zn
0=k∈Zn
Define
∞
1
cμ = .
lμ
l=1
Then cμ < ∞ when μ > 1 and
∞ ∞
1 1 n+m−1
|k|−τ = τ
· #{k ∈ Zn
: |k| = m} ≤ 2 n
C
τ m
m m
0=k∈Zn m=1 m=1
∞
2n−1 1
≤2 τ −n+1
= 22n−1 cτ −n+1 < ∞
m=1
m
when τ > n, here #S denotes the number of the elements of the set S and Csk are
binomial coefficients. (4.5) is verified with
c = 4n cμ cτ −n+1 . (4.6)
Therefore, the lemma is completed.
Remark 4.5. We may define Iλ,μ,τ (ω) to be empty for any resonant frequency ω and
any λ > 0, any μ and any τ > 0 because no number t satisfies (4.4) in this case. It is
possible that the set Iλ,μ,τ (ω) may still be empty even for nonresonant frequencies ω
if here the numbers μ and τ are not properly chosen. Anyway, the above lemma shows
that if μ > 1 and τ > n, then the set Iλ,μ,τ (ω) has positive Lebesgue measure and
hence is nonempty for any λ > 0.
Remark 4.6. If λ1 > λ2 > 0, then Iλ1 ,μ,τ (ω) ⊂ Iλ2 ,μ,τ (ω). Therefore, if ω is a
nonresonant frequency and μ > 1 and τ > n, then the set of all real numbers t
satisfying (4.4) for some λ > 0 has full Lebesgue measure in any measurable set
of R. It should be an interesting number theoretic problem to study the cases when
μ ≤ 1 or τ ≤ n. In numerical analysis, the step sizes are usually considered only in a
bounded interval. We take the interval [−1, 1] as illustration without loss of generality.
Lemma 4.7. For a nonresonant frequency ω = (ω1 , ω2 , · · · , ωn ), assume 0 < λ <
2π
with |ω| = max |ωj |. If −1 ≤ μ ≤ 1 and μ + τ > n + 1, then we have
|ω| 1≤j≤n
meas [−1, 1] \ Iλ,μ,τ (ω) ≤ "
cλ, (4.7)
where "
c is a positive number depending not only on n, μ and τ but also on |ω|.
13.4 Resonant and Diophantine Step Sizes 571
Proof. The set [−1, 1] \ Iλ,μ,τ (ω) is contained in the union of all subintervals
% &
2πl λ 2πl λ
− μ τ, + μ τ
k, ω l |k| k, ω l |k|
2π
where Nl,λ = − λ l which is positive for positive l. We will use the following
|ω|
estimate which is easy to prove:
⎧
1 ⎨ cτ −n+1 , 0 < N ≤ 1,
≤ 1 (4.9)
mτ −n+1 ⎩ , N > 1.
m>N (τ − n)(N − 1)τ −n
Assume lλ is the integer such that Nlλ ,λ ≤ 1 and Nlλ +1,λ > 1. Then (4.7) is verified
with
⎛ ⎞
⎜
lλ ∞
⎟
⎜ 1 1 1 ⎟
"
c = 4n ⎜cτ −n+1 + %% & &τ −n ⎟ (4.10)
⎝ lμ τ −n 2π ⎠
l=1 l=lλ +1 lμ −λ l−1
|ω|
2π
which is finite because the conditions μ + τ > n + 1 and 0 < λ < guarantee the
|ω|
lλ
1
convergence of the infinite summation in (4.1). If lλ = 0, then we take =0
lμ
l=1
and hence the first term in the bracket of Equation (4.10) disappears in this case. Note
that here the number "c depends also on λ, but this dependence is not fatal essentially
2π
because the only harmful case is when λ is close to . However, this case is not of
|ω|
π
interest and may always be avoided. For example, we simply assume 0 < λ ≤ in
|ω|
the lemma. The proof of Lemma 4.7 is completed.
μ + τ > n + 1, which automatically implies that τ > n. One may also consider how
big the set Iλ,μ,τ (ω) is in other unit intervals with integer endpoints, but we do not go
further in this direction.
Remark 4.8. It remains to study the set Iλ,μ,τ (ω) in other cases: μ < −1 or τ ≤ n
or μ + τ ≤ n + 1. I believe the Lebesgue measure of the set is zero in each of these
cases. It is also an interesting problem to calculate the Hausdorff dimensions of the
set Iλ,μ,τ (ω) in all of these cases. The cases when −1 ≤ μ ≤ 1 and ν = n − μ + 1
and when μ > 1 and τ = n should be particularly interesting. In all other cases, I
intend to believe the set is empty. Note that a special case when n = 1, μ = 0 and
τ = n − μ + 1 = 2 with ω = 2π just corresponds to the classical diophantine problem
on approximating an irrational number by rational ones.
To any nonresonant frequency ω in Rn , we have associated a 3-parameter family
of sets Iλ,μ,τ (ω) on the real line. The set Iλ,μ,τ (ω) has positive Lebesgue measure
and hence is nonempty if μ ≥ −1, τ > n, μ + τ > n + 1 and λ > 0 suitably small
(in the case when μ > 1 and τ > n, Iλ,μ,τ (ω) has positive Lebesgue measure for
any λ > 0). But to guarantee an invariant torus of the frequency tω for symplectic
algorithms with the step size t, it seems that the only way is to require tω satisfy a
diophantine condition of the type (1.1) (J. Mather showed in [Mat88] ) that for any exact
area-preserving twist mapping, an invariant circle with any Liouville frequency can be
destroyed by arbitrarily small perturbations in C ∞ -topology). This is the case when
one requires both ω be a diophantine frequency and t be a diophantine step size with
respect to the ω, as the following lemma shows.
Lemma 4.9. Let γ > 0 and 0 < λ ≤ 1. Then for any ω ∈ Ωγ (τ1 )6 and any t ∈
[−1, 1] ∩ Iλ,μ,τ2 (ω), we have
ik,tω |t|"
γ
e − 1 ≥ , 0 = k ∈ Zn , (4.11)
|k|μ+τ1 +τ2
where
2λγ
"=
γ μ . (4.12)
1
π 1+ |ω|
2π
Proof. It is easy to prove that for k ∈ Zn , k = 0, there exists l ∈ Z such that
ik,tω 2
e − 1 ≥ k, tω + 2πl.
π
We have two cases:
1◦ l = 0. Since ω ∈ Ωγ (τ1 ),
ik,tω 2 2|t|γ
e − 1 ≥ k, tω ≥ ;
π π|k|τ1
6
We denote by Ωγ (τ ), the set of all vectors ω ∈ Rn satisfying the diophantine condition of
the form γ
|k, ω| ≥ k ∈ Zn .
, 0=
|k|τ
13.4 Resonant and Diophantine Step Sizes 573
But
π ik,tω
|2πl| ≤ |k, tω + 2πl| + |k, tω| ≤ e − 1 + |t||k, ω|
2
≤ π + |t| |ω||k|,
therefore,
ik,tω 2λγ
e − 1 ≥ % &μ .
1 |t|
π + |ω| |k|μ+τ1 +τ2
2 2π
Combining the two cases, (4.11) is verified and hence Lemma 2.4 is proved.
#
From the above lemmas and the fact that meas Rn \ Ωγ (τ ) = 0 for τ >
γ>0
n − 1, we conclude that for almost all ω ∈ Rn and almost all t ∈ [−1, 1], tω satisfies
a diophantine condition of the mapping type (2.5). As the step size of a difference
scheme, however, t may fall into an arbitrarily small neighbourhood of the origin.
The next lemma shows that for a nonresonant frequency ω ∈ Rn and for μ ≥ −1,
2π
τ > n + 1, μ + τ > n + 1 and 0 < λ < , the set Iλ,μ,τ (ω) has large measure near
|ω|
the origin of the real line.
2π
If λ + δ < , then
|ω|
δ
meas Jλ,μ,τ (ω) ≤ dδ τ −n , (4.13)
where
∞
4n λ 1
d= %% & &τ −n < ∞. (4.14)
τ −n 2π
l=1 lμ −λ l−δ
|ω|
2π
Because 0 < δ + λ < , we have Nlδ > 1. (4.13) follows from (4.9) with the
|ω|
constant d defined by (4.14), which is finite because τ > n and μ + τ − n > 1. (4.15)
is true if, in addition, τ > n + 1.
Proof. For the given ω, we define I(ω) = Iλ,μ,"τ (ω) for some λ > 0, μ > 1 and
τ" > n + 1. By Lemma 3.7, we have for any t ∈ [−1, 1] ∩ I(ω),
|t|"
ik,tω γ
e − 1 ≥ , 0 = k ∈ Zn
|k|μ+τ +"τ
" given by (4.12). The analytic version of Theorem 2.4 may be applied and there-
with γ
fore, for a symplectic algorithm applied to the given system7 , we can find a positive
number δ0 , which depends on the numbers n, γ, τ , λ, μ, τ" and |ω| and on the non-
degeneracy and the analyticity of the system and, of course, also on the algorithm,
such that the algorithm has an invariant torus of the frequency tω with the required
approximating property to the corresponding invariant torus of the system if the step
size t falls into the set [−δ0 , δ0 ] ∩ I(ω). It follows from Lemma 3.8 that the set I(ω)
has density one at the origin because we have chosen μ > 1 and τ" > n + 1.
Remark 4.12. In practical computations, one would like to choose big step sizes. It
is interesting to look at how the δ0 in Theorem 4.11 depends on the nonresonance
property of the frequency ω and how the δ0 relates to the size of the diophantine set
I(ω) of step sizes. It is known that the parameters γ and ν describe the nonresonance
property of the frequency ω and the parameters λ, μ and ν" determine the size of the
set I(ω). Among them, the most interesting are γ and λ because we may fix all others
in advance without loss of generality. For a given ω, we define γ to be the biggest
one such that (4.17) holds for a fixed τ > n − 1. It is easy to see, from Lemma 4.9
2
and Theorem 2.4, that δ0 may be chosen to be proportional to (γλ) s , where s is the
order of accuracy of the algorithm considered in Theorem 4.11. Note that the more
nonresonant the ω is, the bigger γ will be and therefore the bigger δ0 is admitted. On
the other hand, for a given ω, the bigger step size is taken, the bigger λ has to be
chosen and in this case, the set I(ω) turns out to be smaller. But anyway, the set I(ω)
is of density one at the origin. Consequently, to simulate an invariant torus, one has
much more possibilities to select available small step sizes than to select available big
ones.
Remark 4.13. It is interesting to make some comparisons between Theorem 3.4 and
Theorem 4.11. Theorem 3.4 shows that a symplectic algorithm applied to an analytic
nondegenerate integrable Hamiltonian system has so many invariant tori that the tori
form a set of positive Lebesgue measures in the phase space if the step size of the algo-
rithm is sufficiently small and fixed in an arbitrary way. No additional nonresonance or
diophantine condition is imposed on the step size. But the set of frequencies of the in-
variant tori depends on the step size and, therefore, changes in general as the step size
changes. It is a fact that the measure of the set of frequencies of the invariant tori be-
comes larger and larger as the step size gets smaller and smaller. These sets, however,
may not intersect at all for step sizes taken over any interval near the origin. There-
fore, the invariant tori of any frequencies may not be guaranteed for any symplectic
algorithm with step size randomly taken in any neighbourhood of the origin. Theorem
7
So far, the available symplectic algorithms are exact symplectics when they are applied to
global Hamiltonian systems and analytics when applied to analytic systems.
576 13. KAM Theorem of Symplectic Algorithms
4.11 shows that an invariant torus with any fixed diophantine frequency of an analytic
nondegenerate integrable Hamiltonian system can always be simulated very well by
symplectic algorithms for any step size in a Cantor set of positive Lebesgue measure
near the origin. The following theorem shows that one can simulate simultaneously
any finitely many invariant tori of given diophantine frequencies by symplectic algo-
rithms with a sufficiently big probability to select available step sizes. The step sizes,
of course, also have to be restricted to a Cantor set.
Proof. The proof of Theorem 4.14 follows from Theorem 4.11 and
L
N
τ (A ) = (−δ, δ) \ Iλ,μ,"
N N
τ (A )
Iλ,μ," = Iλ,μ,"τ (ω j ), δ
Jλ,μ," N N N
τ (A )
j=1
with given λ > 0, μ ≥ −1, τ" > n + 1 and μ + τ" > n + 1. Then we have
τ" −n
τ (A ) ≤ N dδ
δ N
meas Jλ,μ,"
2π
if λ + δ < , where |AN | = max |ω j | and d is defined by (4.14) where τ is
|AN | 1≤j≤N
replaced by τ" and |ω| replaced by |AN |. Consequently, the set Iλ,μ,"
N N
τ (A ) has density
one at the origin. Moreover, for any t ∈ [−1, 1] ∩ Iλ,μ,"
N N
τ (A ), we have
ik,tωj |t|"
γ
e − 1 ≥ , 0 = k ∈ Zn , j = 1, 2, · · · , N
|k|μ+τ +"τ
Remark 4.16. There have been some works about exponential stability of symplec-
tic algorithms in simulating invariant tori with given diophantine frequencies of in-
tegrable or nearly integrable systems (Benettin and Giorgilli (1994)[BG94] , Hairer and
Lubich in 1997[HL97] and Stoffer in 1998[Sto98b] ). The result, for example, of Hairer and
13.4 Resonant and Diophantine Step Sizes 577
Lubich[HL97] shows that during a very long interval of iteration steps (exponentially
long in 1/t ), the numerical orbits of a symplectic algorithm approximate the exact
orbits of some perturbed Hamiltonian system8 with a very small error (exponentially
small in −1/t ) if the starting values of the numerical orbits and the exact ones are the
same and are taken on the invariant torus of the perturbed system (the invariant torus
is guaranteed by the KAM theorem)[HL97] (Corollary 7) or taken in a neighbourhood
of the invariant torus with the radius of order t2n+2 (this is easily derived from Hairer
and Lubich (1997, Corollary 8)), here n is the degrees of freedom of the Hamiltonian
system and t is the step size of the algorithm which is assumed to be sufficiently small.
Theorems 4.11 and 4.14 show that one may generate quasi-periodic (therefore, perpet-
ually stable) numerical orbits using a symplectic algorithm which approximate exact
quasi-periodic orbit of an analytic nondegenerate integrable Hamiltonian system if the
step sizes of the algorithm fall into a Cantor set of large density near the origin. As the
step size in this Cantor set gets smaller and smaller, more and more stable numerical
orbits appear. For such a stability consideration, Theorem 3.4 shows much more: the
perpetually stable numerical orbits take up a large set of the phase space so that the
Lebesgue measure of the set approaches the Lebesgure measure of the phase space
as the step size approaches zero. Due to the well-known topological confinement of
the phase plane between invariant closed curves, this implies the perpetual stability of
symplectic algorithms applied to one degree of freedom systems for any initial values
if the step size is small.
8
The perturbed Hamiltonian system approximates the symplectic algorithm and is determined
uniquely, in the setting of the backward analysis, by the algorithm and the Hamiltonian
system which the algorithm applies to[Hai94] .
Bibliography
[Fen91] K. Feng: The Hamiltonian Way for Computing Hamiltonian Dynamics. In R. Spigler,
editor, Applied and industrial Mathmatics, pages 17–35. Kluwer, The Netherlands, (1991).
[FQ03] K. Feng and M. Q. Qin: Symplectic Algorithms for Hamiltonian Systems. Zhejiang
Science and Technology Publishing House,Hangzhou, in Chinese, First edition, (2003).
[Gar96] B. M. Garay: On structural stability of ordinary differential equations with respect to
discretization methods. Numer. Math., 72:449–479, (1996).
[Hai94] E. Hairer: Backward analysis of numerical integrators and symplectic methods. Annals
of Numer. Math., 1:107–132, (1994).
[HL97] E. Hairer and Ch. Lubich: The life-span of backward error analysis for numerical
integrators. Numer. Math., 76:441–462, (1997).
[HLW02] E. Hairer, Ch. Lubich, and G. Wanner: Geometric Numerical Integration. Num-
ber 31 in Springer Series in Computational Mathematics. Springer-Verlag, Berlin, (2002).
[HNW93] E. Hairer, S. P. Nørsett, and G. Wanner: Solving Ordinary Differential Equations I,
Nonstiff Problems. Springer-Verlag, Berlin, Second revised edition, (1993).
[HS81] W. H. Hundsdorfer and M. N. Spijker: A note on B-stability of Runge–Kutta methods.
Numer. Math., 36:319–331, (1981).
[HS94] A. R. Humphries and A. M. Stuart: Runge–Kutta methods for dissipative and gradient
dynamical systems. SIAM J. Numer. Anal., 31(5):1452–1485, (1994).
[Kol54b] A. N. Kolmogorov: On conservation of conditionally periodic motions under small
perturbations of the Hamiltonian. Dokl. Akad. Nauk SSSR,, 98:527–530, (1954).
[Laz74] V. F. Lazutkin: On Moser’s theorem on invariant curves. In Voprsoy raspr. seism. voln.
vyp. Nauka Leningrad, 14:105–120, (1974).
[Li99] M. C. Li: Structural stability for Euler method. SIAM J. Math. Anal., 30(4):747–755,
(1999).
[LR05] B. Leimkuhler and S. Reich: Simulating Hamiltonian Dynamics. Cambridge Univer-
sity Press, Cambridge, First edition, (2005).
[Mat88] J. Mather: Destruction of invariant circles. Ergod. Theory & Dynam. Sys, 8:199–214,
(1988).
[Mos62] J. Moser: On invariant curves of area-preserving mappings of an annulus. Nachr.
Akad. Wiss. Gottingen, II. Math.-Phys., pages 1–20, (1962).
[Nei82] A. I. Neishtadt: Estimates in the Kolmogorov theorem on conservation of condition-
ally periodic motions. J. Appl. Math. Mech., 45(6):766–772, (1982).
[Pös82] J. Pöschel: Integrability of Hamiltonian systems on Cantor sets. Comm. Pure and
Appl. Math., 35:653–695, (1982).
[Rüs81] H. Rüssmann: On the existence of invariant curves of twist mappings of an anulus.
In J. Palis, editor, Geometric Dynamics, Lecture Notes in Math. 1007, pages 677–718.
Springer-Verlag, Berlin, (1981).
[Rüs90] H. Rüssmann: On twist Hamiltonian. In in Colloque Internationa: Mécanique céleste
et systèmes hamiltoniens. Marseille, (1990).
[SH96] A.M. Stuart and A.R. Humphries: Dynamical Systems and Numerical Analysis. Cam-
bridge University Press, Cambridge, Second edition, (1996).
[Sha91] Z. J. Shang: On the KAM theorem of symplectic algorithms for Hamiltonian systems,.
Ph.D. thesis (in Chinese), Computing Center, Academia Sinica, (1991).
[Sha99] Z. Shang: KAM theorem of symplectic algorithms for Hamiltonian systems. Numer.
Math., 83:477–496, (1999).
[Sha00a] Z. J. Shang: A note on the KAM theorem for symplectic mappings. J. Dynam.
Differential eqns., 12(2):357–383, (2000).
[Sha00b] Z. J. Shang: Resonant and diophantine step sizes in computing invariant tori of
Hamiltonian systems. Nonlinearity, 13:299–308, (2000).
[SSC94] J. M. Sanz-Serna and M. P. Calvo: Numerical Hamiltonian Problems. AMMC 7.
Chapman & Hall, London, (1994).
[Sto98a] D. Stoffer: On the qualitative behavior of symplectic integrator. II: Integrable systems.
J. of Math. Anal. and Applic., 217:501–520, (1998).
580 Bibliography
t̃ = t, (1.3)
q̃ i = g i (ε, t, q), (1.4)
where
d
g i (ε, t, q) = φi (t, q) := δq i (t). (1.5)
d ε ε=0
In other words, the deformation (1.3) – (1.4) transforms the curve q i (t) into a
family of curves q̃ i (ε, t̃) in Q denoted by Cεba which are determined by
t̃ = t, (1.6)
q̃ i = g i (ε, t, q(t)). (1.7)
Thus, we obtain a (sufficiently small) set of curves Cεba around Cab . Corresponding to
this set of curves there is a set of Lagrangian and action functionals
- b
d i
S(q (t)) −→ S(q̃ (ε, t̃)) =
i i
L(q̃ i (ε, t̃), q̃ (ε, t̃)) d t̃. (1.8)
a d t̃
For the fixed endpoints, φi (a, q(a)) = φi (b, q(b)) = 0, the requirement of Hamilton’s
principle, δS = 0, yields the Euler–Lagrange equation for q(t)
14.1 Total Variation in Lagrangian Formalism 583
∂L d ∂L
− = 0. (1.10)
∂ qi d t ∂ q̇ i
ωL := dθL . (1.12)
∂ ∂
V = ξ(t, q) + φi (t, q) i , (1.15)
∂t ∂q
where
t̃ = f (ε, t, q), q̃ i = g i (ε, t, q) (1.17)
with
d d
f (ε, t, q) = ξ(t, q) := δt, g i (ε, t, q) = φi (t, q) := δq i . (1.18)
dε ε=0 dε ε=0
The deformations (1.17) transform a curve q i (t) in Q denoted by Cab into a set of
curves q̃ i (ε, t̃) in Q denoted by Cεb̃ã , determined by
584 14. Lee-Variational Integrator
Before calculating the total variation of S, we will introduce the first-order prolonga-
tion of V denoted as pr1 V
∂ ∂ ∂
pr1 V = ξ(t, q) + φi (t, q) i + αi (t, q, q̇) i , (1.20)
∂t ∂q ∂ q̇
here
αi (t, q, q̇) = Dt φi (t, q) − q̇ i Dt ξ(t, q), (1.21)
where Dt denotes the total derivative with respect to t, for example,
∂φk
Dt φk (t, q i ) = φkt + φkqi q̇ i , φkt = .
∂t
For prolongations of the vector field and the related formulae, refer to[Olv93] .
Now, we let us calculate the total variation of S straightforwardly:
- b̃
d i d d
δS = S(q̃ (ε, t̃)) = L t̃, q̃ i (ε, t̃), q̃ i (ε, t̃) d t̃
dε ε=0 dε ε=0 ã dt̃
- b
d d dt̃
= L t̃, q̃ i (ε, t̃), q̃ i (ε, t̃) d t t̃ = f (ε, t, q(t))
dε ε=0 a d t̃ dt
- b - b
d d i
= L t̃, q̃ i (ε, t̃), q̃ (ε, t̃) d t + L t, q i (t), q̇ i (t) Dt ξ d t
a d ε ε=0 d t̃ a
- b - b
∂L ∂L ∂L
= ξ + i φi + i (Dt φi − q̇ i Dt ξ) d t + LDt ξ d t
a ∂t ∂q ∂ q̇ a
- b !
∂L d ∂L i ∂L d ∂L i
= + q̇ − L ξ + − φ dt
a ∂t dt ∂ q̇ i ∂q i dt ∂ q̇ i
!
∂L ∂L b
+ L − i q̇ i ξ + i φi . (1.22)
∂ q̇ ∂ q̇ a
d d t̃ d d
= t̃ = Dt ξ.
d ε ε=0 d t d t d ε ε=0
If ξ(a, q(a)) = ξ(b, q(b)) = 0 and φi (a, q(a)) = φi (b, q(b)) = 0, the requirement
of δS = 0 yields the equation from ξ, the variation along the base manifold, i.e., the
time.
∂L d ∂L i
+ q̇ − L = 0, (1.23)
∂t dt ∂ q̇ i
and the Euler–Lagrange equation from φi , the variation along the fiber, i.e., the con-
figuration space,
∂L d ∂L
− = 0. (1.24)
∂q i dt ∂ q̇ i
14.1 Total Variation in Lagrangian Formalism 585
d ∂L
H = 0, H := i
q̇ i
− L . (1.25)
dt ∂ q̇
By expanding the left-hand side of (1.25), we obtain
% & % &
d ∂L i ∂L d ∂L
q̇ − L = − − q̇ i . (1.26)
dt ∂ q̇ i ∂q i dt ∂ q̇ i
we can define the extended Lagrangian 1-form on Q(1) from the second term in (1.22)
% &
∂L i ∂L
ϑL := L − i
q̇ dt + d qi . (1.29)
∂ q̇ ∂ q̇ i
Suppose g i (t, vqi ) is a solution of (1.24) depending on the initial condition vqi ∈ Q(1) .
Restricting q̃ i (ε, t̃) to the solution space of (1.24) and using the same method in[MPS98] ,
it can be proved that the extended symplectic 2-form is preserved:
(pr1 g i )∗ ΩL = ΩL ,
ΩL := dϑL , (1.30)
d i
where pr1 g i (s, vqi ) = s, g i (s, vqi ), g (s, vqi ) denotes the first-order prolonga-
ds
tion of g i (s, vqi )[Olv93] .
(tk+1 − tk )D2 L(tk , qk , tk+1 , qk+1 ) + (tk − tk−1 )D4 L(tk−1 , qk−1 , tk , qk ) = 0,
(1.35)
and
for all k ∈ {1, 2, · · · , N − 1}. Here Di denotes the partial derivative of L with respect
to the i-th argument. The Equation (1.35) is the discrete Euler–Lagrange equation.
The Equation (1.36) is the discrete energy conservation law for a conservative L. The
integrator (1.35) – (1.36) is the Kane–Marsden–Ortiz integrator.
Using the discrete flow Φ, the Equations (1.35) and (1.36) become
respectively. If (tk+1 − tk )D2 L and (tk+1 − tk )D1 L − L are invertible, the Equations
(1.37) and (1.38) determine the discrete flow Φ under the consistency condition
((tk+1 −tk )D1 L−L)−1 ◦(D3 L+L) = ((tk+1 −tk )D2 L)−1 ◦(tk −tk−1 )D4 L. (1.39)
Now, we will prove that the discrete flow Φ preserves a discrete version of the
extended Lagrange 2-form ΩL . As in continuous case, we will calculate d S for varia-
tions with variable end points.
N −1
N
+ L(tk , qk , tk+1 , qk+1 )(−δtk ) + L(tk−1 , qk−1 , tk , qk )δtk
k=0 k=1
N −1
= (D2 L(tk , qk , tk+1 , qk+1 )(tk+1 − tk )
k=1
+D4 L(tk−1 , qk−1 , tk , qk )(tk − tk−1 ))δqk
N −1
+ (D1 L(tk , qk , tk+1 , qk+1 )(tk+1 − tk )
k=1
+D3 L(tk−1 , qk−1 , tk , qk )(tk − tk−1 )) (1.40)
+L(tk−1 , qk−1 , tk , qk ) − L(tk , qk , tk+1 , qk+1 ))δtk
+D2 L(t0 , q0 , t1 , q1 )(t1 − t0 )δq0 + D4 L(tN −1 , qN −1 , tN , qN )(tN − tN −1 )δqN
+(D1 L(t0 , q0 , t1 , q1 )(t1 − t0 ) − L(t0 , q0 , t1 , q1 ))δt0
+(D3 L(tN −1 , qN −1 , tN , qN )(tN − tN −1 ) + L(tN −1 , qN −1 , tN , qN ))δtN .
We can see that the last four terms in (1.40) come from the boundary variations. Based
on the boundary variations, we can define two 1-forms on Q × Q
and
having employed the notations in [MPS98] . We regard the pair (θL− , θL+ ) as the discrete
version of the extended Lagrange 1-form ϑL defined in (1.29).
Now, we parameterize the solutions of the discrete variational principle by the
initial condition (t0 , q0 , t1 , q1 ) and restrict S to that solution space. Then Equation
(1.40) becomes
Finally, we have shown that the discrete flow Φ preserves the discrete extended La-
grange 2-form ΩL
Φ∗ (ΩL ) = ΩL . (1.49)
Now, the variational integrator (1.35), the discrete energy conservation law (1.36),
and the discrete extended Lagrange 2-form ΩL converge to their continuous counter-
parts as tk+1 → tk , tk−1 → tk .
Consider a conservative Lagrangian L(q, q̇). For simplicity, we choose the discrete
Lagrangian as % &
qk+1 − qk
L(tk , qk , tk+1 , qk+1 ) = L qk , . (1.50)
tk+1 − tk
The variational integrator (1.35) becomes
% &
∂L 1 ∂L ∂L
(qk , Δt qk ) − (qk , Δt qk ) − (qk−1 , Δt qk−1 ) = 0,
∂ qk tk+1 − tk ∂ Δ t qk ∂ Δt qk−1
(1.51)
q − qk q − qk−1
where Δt qk = k+1 , Δt qk−1 = k .
tk+1 − tk tk − tk−1
It is easy to see that, as tk+1 → tk , tk−1 → tk , the Equation (1.51) converges to
∂L d ∂L
− = 0. (1.52)
∂ qk d t ∂ q̇k
The discrete energy conservation law (1.36) becomes
Ek+1 − Ek
= 0, (1.53)
tk+1 − tk
where
% &
∂L qk+1 − qk
Ek+1 = Δt qk − L q k , ,
∂Δt qk tk+1 − tk
% &
∂L qk − qk−1
Ek = Δt qk−1 − L qk−1 , .
∂Δt qk−1 tk − tk−1
The Equation (1.53) converges to
590 14. Lee-Variational Integrator
% &
d ∂L
q̇k − L = 0 (1.54)
dt ∂ q̇k
as tk+1 → tk , tk−1 → tk .
Now, we will consider the discrete extended Lagrange 2-form ΩL defined by
(1.48). By discretization of (1.50), the discrete extended Lagrange 1-form θL+ defined
in (1.42) becomes
% &
+ ∂L ∂L
θL = L(qk , Δt qk ) − Δt qk d tk+1 + d qk+1 . (1.55)
∂ Δ t qk ∂ Δt q k
From (1.55), we can deduce that θL+ converges to the continuous Lagrangian 1-form
ϑL defined by (1.29) as tk+1 → tk , tk−1 → tk . Thus, we obtain
ΩL = dθL+ −→ dϑL = ΩL , tk+1 → tk , tk−1 → tk . (1.56)
In general, the variational integrator (1.35) with fixed time steps does not ex-
actly conserve the discrete energy, and the computed energy will not have secular
variation[GM88,SSC94] . In some cases, such as in discrete mechanics proposed by Lee
in [Lee82,Lee87] , the integrator (1.35) is required to conserve the discrete energy (1.36)
by varying the time steps. In other words, the steps can be chosen according to (1.36)
so that the integrator (1.35) conserves the discrete energy (1.36). The resulting inte-
grator also conserves the discrete extended Lagrange 2-form dθL+ . This fact had not
been discussed in Lee’s discrete mechanics.
Example 1.3. Let us consider an example. For the classical Lagrangian
1
L(t, q, q̇) = q̇ 2 − V (q), (1.57)
2
we choose the discrete Lagrangian L(tk , qk , tk+1 , qk+1 ) as
% &2 % &
1 qk+1 − qk qk+1 − qk
L(tk , qk , tk+1 , qk+1 ) = −V . (1.58)
2 tk+1 − tk 2
The discrete Euler–Lagrange equation (1.35) becomes
% &
qk+1 − qk qk − qk−1 V (q̄k )(tk+1 − tk ) + V (q̄k−1 )(tk − tk−1 )
− + = 0,
tk+1 − tk tk − tk−1 2
(1.59)
which preserves the Lagrange 2-form
% &
1 t − tk
+ k+1 V (q̄k ) d qk+1 ∧ d qk , (1.60)
tk+1 − tk 4
q +q q +q
where q̄k = k k+1
, q̄k−1 = k−1 k
.
2 2
If we take fixed variables tk+1 − tk = tk − tk−1 = h, then (1.59) becomes
qk+1 − 2qk + qk−1 V (q̄k ) + V (q̄k−1 )
2
+ = 0,
h 2
which preserves the Lagrange 2-form
% &
1 h
+ V (q̄k ) d qk+1 ∧ d qk .
h 4
14.2 Total Variation in Hamiltonian Formalism 591
The variational principle in Hamiltonian formalism seeks the curves (q i (t), pi (t))
for which the action functional S is stationary under variations of (q i (t), pi (t)) with
fixed end points. We will first define the variation of (q i (t), pi (t)).
Let
n
∂
n
∂
V = φi (qq , p ) i + ψ i (qq , p ) i , (2.2)
∂q ∂p
i=1 i=1
q̃ i = f i (ε, q , p ), (2.3)
p̃i = g i (ε, q , p ), (2.4)
Let (q i (t), pi (t)) be a curve in T ∗ Q. The transformation (2.3) and (2.4) transforms
(q i (t), pi (t)) into a family of curves
i
q̃ (t), p̃i (t) = f i (ε, q (t), p (t)), g i (ε, q (t), p (t)) .
Next, we will define the variation of q i (t), pi (t) :
d
δ q i (t), pi (t) =: q̃ i (t), p̃i (t) = φi (qq , p ), ψ i (qq , p ) . (2.5)
dε ε=0
Next, we will calculate the variation of S at q i (t), pi (t) as follows:
d
δS = S (q̃ i (t), p̃i (t))
d ε ε=0
d i i
= S (f (ε, q (t), p (t)), g (ε, q (t), p (t)))
dε ε=0
- b% d
d
= g i ε, q (t), p (t) f i ε, q (t), p (t)
dε ε=0 a dt
&
−H f i (ε, q (t), p (t)), g i (ε, q (t), p (t)) d t
- b % & % & ! b
i ∂H i i ∂H
= q̇ − i
ψ + − ṗ − i
φi d t + pi φi .
a ∂p ∂q a
(2.6)
i
If φi q (a), p (a)
i φ qi(b),p (b) = 0, the requirement of δS = 0 yields the Hamil-
ton equation for q (t), p (t) :
14.2 Total Variation in Hamiltonian Formalism 593
∂H ∂H
q̇ i = , ṗi = − . (2.7)
∂ pi ∂ qi
If we drop the requirement of φi q (a), p (a) φi q (b), p (b) = 0, we can naturally
∗ i i
i oni T Q
obtain the canonical 1-form from the second term in (2.6): θ = p dq . Fur-
thermore, restricting (q̃ (t), p̃ (t)) to the solution space of (2.7), we can prove that
the solution of (2.7) preserves the canonical 2-form ω = d θL = d pi ∧ d q i .
On the other hand, it is not necessary to restrict ((q̃ i (t), p̃i (t)) to the solution space
of (2.7). Introducing the Euler–Lagrange 1-form
∂H ∂H
E(q i , pi ) = q̇ i − i
d pi + − ṗi − i
d qi , (2.8)
∂p ∂q
namely, the necessary and sufficient condition for symplectic structure preserving is
that the Euler–Lagrange 1-form (2.8) is closed[GLW01a,GLWW01,GLW01b,GW03] .
Based on the above-given variational principle in Hamiltonian formalism and
using the ideas of discrete Lagrange mechanics[Ves88,Ves91b,MPS98,WM97] , we can de-
velop a natural version of discrete Hamilton mechanics with fixed time steps and
derive symplectic integrators for Hamilton canonical equations from a variational
perspective[GLWW01] .
However, the symplectic integrators obtained in this way are not energy-preserving,
in general, because of its fixed time steps[GM88] . An energy-preserving symplectic in-
tegrator is a more preferable and natural candidate of approximations for conservative
Hamilton equations since the solution of conservative Hamilton equations is not only
symplectic but also energy-preserving. To attain this goal, we use variable time steps
and a discrete total variation calculus developed in [Lee82,Lee87,KMO99,CGW03] . The basic
idea is to construct a discrete action functional with variable time steps and then apply
a discrete total variation calculus. In this way, we can derive symplectic integrators
and their associated energy conservation laws. These variationally derived symplectic
integrators are two-step integrators. If we take fixed time steps, the resulting integra-
tors are equivalent to the symplectic integrators derived directly from the Hamiltonian
systems in some special cases.
∂ ∂ ∂
V = ξ(t, q , p ) + φi (t, q , p ) i + ψ i (t, q , p ) i . (2.10)
∂t ∂q ∂p
594 14. Lee-Variational Integrator
t̃ = h(ε, t, q , p ), (2.11)
q̃ i = f i (ε, t, q , p ), (2.12)
p̃i = g i (ε, t, q , p ), (2.13)
where
d
h(ε, t, q , p ) = ξ(t, q , p ), (2.14)
d ε ε=0
d
f i (ε, t, q , p) = φi (t, q , p), (2.15)
d ε ε=0
d
g i (ε, t, q , p ) = ψ i (t, q , p ). (2.16)
d ε ε=0
The transformation (2.11) – (2.13) transforms a curve (q i (t), pi (t)) into a family of
curves (q̃ i (ε, t̃), p̃i (ε, t̃)) determined by
t̃ = h ε, t, q (t), p (t) , (2.17)
i i
q̃ = f ε, t, q (t), p (t) , (2.18)
i i
p̃ = g ε, t, q (t), p (t) . (2.19)
q̃ i (ε, t̃) = f i (ε, h−1 (ε, t̃), q (h−1 (ε, t̃)), p (h−1 (ε, t̃))), (2.20)
q̃ i (ε, t̃) = f i (ε, h−1 (ε, t̃), q (h−1 (ε, t̃)), p (h−1 (ε, t̃))). (2.21)
Before calculating the variation of S directly, we will first consider the first-order
prolongation of V ,
1 ∂ i ∂ i ∂ i ∂ i ∂
pr V = ξ(t, q , p ) + φ (t, q , p ) i + ψ (t, q , p ) i + α (t, q , p , ·q
·q, ·p
·p) i + β (t, q , p , q̇ , ṗ ) i ,
∂t ∂q ∂p ∂ q̇ ∂ ṗ
(2.22)
where pr1 V denotes the first-order prolongation of V and
For prolongation of vector field and formulae (2.23) and (2.24), refer to[Olv93] .
Now, let us calculate the variation of S directly as follows:
d
δS = S (q̃ i (ε, t̃), p̃i (ε, t̃))
d ε ε=0
- b̃
d i d i
= p̃ (ε, t̃) q̃ (ε, t̃) − H q̃ i (ε, t̃), p̃i (ε, t̃) d t̃
d ε ε=0 ã dt̃
14.2 Total Variation in Hamiltonian Formalism 595
-
d b d d t̃
= p̃i (ε, t̃) q̃ i (ε, t̃) − H (q̃ i (ε, t̃), p̃i (ε, t̃) dt
d ε ε=0 d t̃ dt
a
t̃ = h ε, t, q (t), p (t)
- b
d d
= p̃i (ε, t̃) q̃ i (ε, t̃) − H q̃ i (ε, t̃), p̃i (ε, t̃) d t
a d ε ε=0 d t̃
- b
i
+ p (t)q̇ i (t) − H q i (t), pi (t) Dt ξ d t (2.25)
a
- d
b ∂ H i i ∂ H i
= H q i (t), pi (t) ξ + − ṗi − φ + q̇ − ψ dt
a dt ∂ qi ∂ pi
@ Ab
+ pi φi − H(q i , pi )ξ . a
(2.26)
In (2.25), we have used (2.14)and the fact
d d t̃ d d
= t̃ = Dt ξ.
d ε ε=0 d t d t d ε ε=0
In (2.26),
we have used theprolongation formula (2.23).
If ξ a, q (a), p (a) = ξ b, q (b), p (b) = 0 and φi a, q (a), p (a) = φi b, q (b), p (b)
= 0, the requirement of δS = 0 yields the Hamilton canonical equation
∂H ∂H
q̇ i = , ṗi = − (2.27)
∂pi ∂q i
N −1
S= L(tk , q k , p k , tk+1 , q k+1 , p k+1 )(tk+1 − tk ), (2.31)
k=0
Here, (tk+1 , q k+1 , p k+1 ) for all k ∈ (1, 2, · · · , N − 1) are found from the following
discrete Hamilton canonical equation
(tk+1 − tk )D2 L(tk , q k , p k , tk+1 , q k+1 , p k+1 ) + (tk − tk−1 )D5 L(tk−1 , q k−1 , p k−1 , tk , q k , p k ) = 0,
(2.33)
(tk+1 − tk )D3 L(tk , q k , p k , tk+1 , q k+1 , p k+1 ) + (tk − tk−1 )D6 L(tk−1 , q k−1 , p k−1 , tk , q k , p k ) = 0
we call (2.33) and (2.34) a symplectic-energy integrator. We will do this directly from
the variational point of view, consistent with the continuous case[MPS98] .
As in the continuous case, we will calculate dS for variations with varied end
points.
dS(t0 , q 0 , p 0 , · · · , tN , q N , p N ) · (δt0 , δqq 0 , δpp0 , · · · , δtN , δqq N , δppN )
N −1 % &
= D2 L(vv k )δqq k + D5 L(vv k )δqq k+1 + D3 L(vv k )δppk + D6 L(vv k )δppk+1 (tk+1 − tk )
k=0
N −1
N −1
+ D1 L(vv k )δtk + D4 L(vv k )δtk+1 (tk+1 − tk ) + L(vv k )(δtk+1 − δtk )
k=0 k=0
N −1 % &
= D2 L(vv k )(tk+1 − tk ) + D5 L(vv k−1 )(tk − tk−1 δqq k
k=1
N −1 % &
+ D3 L(vv k )(tk+1 − tk ) + D6 L(vv k−1 )(tk − tk−1 δppk
k=1
N −1
+ D1 L(vv k )(tk+1 − tk ) + D4 L(vv k−1 )(tk − tk−1 ) + L(vv k−1 ) − L(vv k ) δtk
k=1
+D2 L(vv 0 )(t1 − t0 )δqq 0 + D3 L(vv 0 )(t1 − t0 )δp0 + D1 Lvv 0 )(t1 − to ) − L(vv 0 ) δt0
+D5 L(vv N −1 )(tN − tN −1 )δqq N + D6 L(vv N −1 )(tN − tN −1 )δppN
+ D4 L(vv N −1 )(tN − tN −1 ) − L(vv N −1 ) δtN ,
(2.35)
where v k = (tk , q k , p k , tk+1 , q k+1 , p k+1 ) (k = 0, 1, · · · , N − 1). We can see that the
last six terms in (2.35) come from the boundary variations. Based on the boundary
variations, we can define two 1-forms on P × P ,
θL− (vv k ) = D2 L(vv k )(tk+1 − tk )dqq k + D3 L(vv k )(tk+1 − tk )dppk
+D1 L(vv k )(tk+1 − tk ) − L(vv k )dtk (2.36)
and
θL+ (vv k ) = D5 L(vv k )(tk+1 − tk )dqq k+1 + D6 L(vv k )(tk+1 − tk )dppk+1
+D4 L(vv k )(tk+1 − tk ) − L(vv k )dtk+1 . (2.37)
Here, we have used the notation in [MPS98] . We regard the pair (θL− , θL+ ) as being the
discrete version of the extended canonical 1-form θ defined in (2.29).
Now, we will parametrize the solutions of the discrete variational principle by
(t0 , q0 , t1 , q1 ), and restrict S to that solution space. Then, Equation (2.35) becomes
d S(t0 , q 0 , p 0 , · · · , tN , q N , p N ) · (δt0 , δqq 0 , δpp0 , · · · , δtN , δqq N , δppN )
= θL− (t0 , q 0 , p 0 , t1 , q 1 , p 1 ) · (δt0 , δqq 0 , δpp0 , δt1 , δqq 1 , δpp1 )
+θL+ (tN −1 , q N −1 , p N −1 , tN , q N , p N ) · (δtN −1 , δqq N −1 , δppN −1 , δtN , δqq N , δppN )
= θL− (t0 , q 0 , p 0 , t1 , q 1 , p 1 ) · (δt0 , δqq 0 , δpp0 , δt1 , δqq 1 , δpp1 )
+(ΦN −1 )∗ θL+ (t0 , q 0 , p 0 , t1 , q 1 , p 1 ) · (δt0 , δqq 0 , δpp0 , δt1 , δqq 1 , δpp1 ). (2.38)
598 14. Lee-Variational Integrator
Finally, we have shown that the discrete flow Φ preserves the discrete extended canon-
ical 2-form ωL :
Φ∗ (ωL ) = ωL . (2.44)
We can now call the coupled difference system (2.33) and (2.34) a symplectic-
energy integrator in the sense that it satisfies the discrete energy conservation law
(2.34) and preserves the discrete extended canonical 2-form ωL .
To illustrate the above-mentioned discrete total variation calculus, we present an
example. We choose L in (2.31) as
q k+1 − q k
L(tk , q k , p k , tk+1 , q k+1 , p k+1 ) = p k+1/2 − H(qq k+1/2 , p k+1/2 ), (2.45)
tk+1 − tk
where
p k + p k+1 q + q k+1
p k+1/2 = , q k+1/2 = k .
2 2
Using (2.33), we can obtain the corresponding discrete Hamilton equation
* +
q k+1 − q k−1 1 ∂H ∂H
− (tk+1 − tk ) (qq ,p ) + (tk − tk−1 ) (qq ,p = 0,
2 2 ∂pp k+1/2 k+1/2 ∂pp k+1/2 k+1/2
* +
p k+1 − p k−1 1 ∂H ∂H
+ (tk+1 − tk ) (qq k+1/2 , p k+1/2 ) + (tk − tk−1 ) (qq k+1/2 , p k+1/2 ) = 0,
2 2 ∂qq ∂qq
(2.46)
p +p
p q +qq
where p k−1/2 = k 2 k−1 , q k−1/2 = k 2 k−1 .
Using (2.34), we can obtain the corresponding discrete energy conservation law
H q k+1/2 , p k+1/2 = H q k+1/2 , p k+1/2 . (2.47)
The symplectic-energy integrator (2.46) and (2.47) preserves the discrete 2-form:
14.2 Total Variation in Hamiltonian Formalism 599
1 d tk + dtk+1
d p k ∧ dqq k+1 + dppk+1 ∧ dqq k − H q k+1/2 , p k+1/2 ∧ . (2.48)
2 2
If we take fixed time steps tk+1 − tk = h (h is a constant), then (2.46) becomes
q k+1 − q k−1 1 ∂H ∂H
= q k+1/2 , p k+1/2 + q k−1/2 , p k−1/2 ,
2h 2 ∂p ∂p
p k+1 − p k−1 1 ∂H ∂H (2.49)
=− q k+1/2 , p k+1/2 ) + (qq k−1/2 , p k−1/2 .
2h 2 ∂q ∂q
Now, we will explore the relationship between (2.49) and the midpoint integrator for
the Hamiltonian system
∂H
q̇q = ,
∂pp
∂H (2.50)
ṗp = − .
∂qq
The midpoint symplectic integrator for (2.50) is
q k+1 − q k ∂H
= q ,p ,
h ∂ p k+1/2 k+1/2
p k+1 − p k ∂H (2.51)
= − q ,p .
h ∂qq k+1/2 k+1/2
The integrator (2.54) is a two-step integrator which preserves dpk ∧dqk+1 . In this case,
we cannot find an one-step integrator which is equivalent to (2.54). In conclusion,
using discrete total variation calculus, we have derived two-step symplectic-energy
integrators. When taking fixed time steps, some of them are equivalent to one-step
integrators derived directly from the Hamiltonian system while the others do not have
this equivalence.
600 14. Lee-Variational Integrator
where
O −I
z = (pp, q )T , J= .
I O
Let us first recall the generating function with normal Darboux matrix of a symplectic
transformation. For details, see Chapters 5 and 6, or [Fen86,FWQW89] .
Suppose α is a 4n × 4n nonsingular matrix with the form
* +
A B
α= ,
C D
σα : Sp(2n) −→ Sm(2n),
σα = (AS + B)(CS + D)−1 = M, for S ∈ Sp(2n), det (CS + D) = 0
where Sp(2n) is the group of symplectic matrices and Sm(2n) the set of symmetric
matrices.
14.2 Total Variation in Hamiltonian Formalism 601
w ) = ∇φ(w
f (w w ), (2.57)
where ∇φ(w w ), · · · , φw2n (w
w ) = φw1 (w w ) and w = (w1 , w2 , · · · , w2n ).
Then, we have
σα : SpD2n −→ Symm(2n),
σα = (A ◦ g + B) ◦ (C ◦ g + D)−1 = ∇φ, for g ∈ SpD2n , det(Cgz + D) = 0
or alternatively
Ag(zz ) + Bzz = (∇φ)(Cg(zz ) + Dzz ),
where ◦ denotes the composition of transformation and the 2n × 2n constant matrices
A, B, C and D are regarded as linear transformations. gz denotes the Jacobian of
symplectic transformation g.
Let φ be the generating function of Darboux type α for symplectic transformation
g.
Conversely, we have
or alternatively
A1 ∇φ(w
w ) + B1w = g(C1 ∇φ(w
w ) + D1w ),
where g is called the symplectic transformation of Darboux type α for the generating
function φ.
For the study of integrators, we will restrict ourselves to normal Darboux matrices,
i.e., those satisfying A + B = 0, C + D = I2n . The normal Darboux matrices can be
characterized as
* +
J2n −J2n 1
α= , E = (I2n + J2n F ), F T = F, (2.58)
E I2n − E 2
and * +
(E − I2n )J2n I2n
α−1 = . (2.59)
EJ2n I2n
* +
J2n −J2n
α= 1 1 . (2.60)
I2n I2n
2 2
Now, we will consider the generating function of the flow of (2.55) and denote it by
etH . The generating function φ(w
w , t) for the flow etH of Darboux type (2.60) is given
by
1 1 −1
∇φ = (J2n ◦ etH − J2n ) ◦ etH + I2n , for small |t|, (2.61)
2 2
w , t) satisfies the Hamilton–Jacobi equation
where φ(w
∂ 1
w , t) = −H w + J2n ∇φ(w
φ(w w , t) (2.62)
∂t 2
and can be expressed by Taylor series in t,
∞
w , t) =
φ(w φk (w)tk , for small |t|. (2.63)
k=1
From (2.61), we can see that the phase flow z := etH z satisfies
z − z ∞
z + z
z − z ) = ∇φ
J2n ( = tj ∇φj . (2.65)
2 j=1
2
where tk−1 , qk−1 , pk−1 and tk , qk , pk are given and tk+1 , qk+1 , pk+1 are unknowns.
In the following numerical experiment, we will use a robust optimization method
suggested in [KMO99] to solve (2.72). Concretely, let
qk+1 −qk−1 1
A= 2
− 2
(tk+1 − tk )pk+1/2 + (tk − tk−1 )pk−1/2 ,
pk+1 −pk−1 3 3
B= 2
+ 12 (tk+1 − tk )(2qk+1/2 − qk+1/2 ) + (tk − tk−1 )(2qk−1/2 − qk−1/2 ) ,
C = 12 p2k+1/2 + 12 (qk+1/2
4 2
− qk+1/2 ) − 12 p2k−1/2 − 12 (qk−1/2
4 2
− qk−1/2 ).
Then, we will minimize the quantity
F = A2 + B 2 + C 2 (2.73)
604 14. Lee-Variational Integrator
Fig. 2.1. The orbits calculated by (2.72), (2.74) left plot q0 = 0.77, p0 = 0 and right plot q0 = 0.99, p0 = 0
Fig. 2.2. The energy evaluation by (2.72), (2.74) left plot q0 = 0.77, p0 = 0 and right plot q0 = 0.99, p0 = 0
over qk+1 , pk+1 and tk+1 under the constraint tk+1 > tk . This constraint guarantees
that no singularities occur in choosing time steps.
We will compare (2.72) with the following integrator with fixed time steps:
qk+1 − qk−1 1
− (pk+1/2 + pk−1/2 ) = 0,
2h 2
pk+1 − pk−1 1 3 3
(2.74)
+ (2qk+1/2 − qk+1/2 ) + (2qk−1/2 − qk−1/2 ) = 0.
2h 2
the Hamiltonian setting and obtain the symplectic-energy integrators. The comprehen-
sive implementation of the obtained integrators is not the subject of present and will
be a topic for future research.
the intervals Ik = [tk , tk+1 ] are called elements. hk = tk+1 − tk .Vh ([a, b]) consists
of piecewise linear function interpolating q(t) at (tk , qk )(k = 0, 1, · · · , N ). Now, we
will derive the expressions of qh (t) ∈ Vh ([a, b]). First, we will construct the basis
functions ϕk (t), which are piecewise linear functions on [a, b] satisfying ϕk (ti ) =
δki (i, k = 0, 1, · · · , N ).
t − t0
1− , t0 ≤ t ≤ t1 ;
ϕ0 (t) = h0
0, otherwise;
t − tN (3.1)
1+ , tN −1 ≤ t ≤ tN ;
ϕN (t) = hN −1
0, otherwise;
and for k = 1, 2, · · · , N − 1,
⎧ t − tk
⎪
⎪ tk−1 ≤ t ≤ tk ;
⎨ 1+ hk−1
,
ϕk (t) = t − tk (3.2)
⎪ 1− , tk ≤ t ≤ tk+1 ;
⎪
⎩ hk
0, otherwise.
N
qh (t) = qk ϕk (t). (3.3)
k=0
−1 - tk+1
* +
N
N
d
N
= L t, (qi ϕi (t), (qi ϕi (t)) d t
tk i=0
d t i=0
k=0
N −1
= L(tk , qk , tk+1 , qk+1 )(tk+1 − tk ), (3.4)
k=0
where
- * +
1 tk+1
N
d
N
L(tk , qk , tk+1 , qk+1 ) = L t, (qi ϕi (t), (qi ϕi (t)) d t
tk+1 − tk tk i=0
d t i=0
- tk+1 * k+1
+
d
k+1
1
= L t, (qi ϕi (t), (qi ϕi (t)) d t.
tk+1 − tk tk dt
i=k i=k
(3.5)
Therefore, restricting to the subspace Vh ([a, b]) of C 2 ([a, b]), the original varia-
tional problem reduces to the extremum problem of the function (3.4) in qk (k =
0, 1, · · · , N ). Note that (3.4) is one of the discrete actions (1.33). Thus, what remains
to be done is just to perform the same calculation on (3.4) as on (1.33). We can then
obtain the discrete Euler–Lagrange equation (1.35) which preserves the discrete La-
grange 2-form (1.48). Therefore, discrete mechanics based on finite element methods
consists of two steps: first, use finite element methods to obtain a kind of discrete
Lagrangian, second, use the method of Veselov mechanics to obtain the variational
integrators.
Let us consider the previous example again. For the classical Lagrangian (1.57),
we choose the discrete Lagrangian L(tk , qk , tk+1 , qk+1 ) as
L(tk , qk , tk+1 , qk+1 )
⎛ * +2 * N +⎞
- tk+1 N
1 ⎝ 1 d
= (qi ϕi (t)) −V (qi ϕi (t)) ⎠ d t
tk+1 − tk tk 2 dt i=0 i=0
- tk+1 * % &2 % &+
1 1 qk+1 − qk tk+1 − t t − tk
= −V qk + qk+1 dt
tk+1 − tk tk 2 tk+1 − tk tk+1 − tk tk+1 − tk
% &2
1 qk+1 − qk
= − F (qk , qk+1 ), (3.6)
2 tk+1 − tk
where
- % &
1 tk+1
tk+1 − t t − tk
F (qk , qk+1 ) = V qk + qk+1 d t. (3.7)
tk+1 − tk tk tk+1 − tk tk+1 − tk
d2 q ∂ V (q)
+ = 0, (3.10)
d t2 ∂q
N
where · is the L2 norm. qh (t) = qk , h = max{hk } and C is a constant
k
k=0
independent of h.
If we use midpoint numerical integration formula in (3.7), we obtain
% &
1 tk+1 tk+1 − t t − tk
F (qk , qk+1 ) = V q k + q k+1 d t
tk+1 − tk tk tk+1 − tk tk+1 − tk
q + qk+1
≈v k .
2
In this case, (3.8) is the same as (1.59). We can also use trapezoid formula or Simpson
formula and so on to integrate (3.7) numerically and obtain another kind of discrete
Lagrangian.
and for k = 1, 2, · · · , N − 1,
⎧ % &% &
⎪
⎪ 2(tk − t) tk − t
⎪
⎪ −1 − 1 , tk−1 ≤ t ≤ tk ;
⎪
⎨ % hk−1 hk−1
&
φk (t) = 2(t − tk ) t − tk (3.14)
⎪
⎪ − 1 − 1 , tk ≤ t ≤ tk+1 ;
⎪
⎪ hk hk
⎪
⎩
0, otherwise;
and for k = 0, 1, · · · , N − 1,
t − tk t − tk
4 1− , tk ≤ t ≤ tk+1 ;
φk+ 12 (t) = hk hk (3.15)
0, otherwise.
Using these basis functions, we will construct subspace Vh2 ([a, b]) of C 2 ([a, b]):
N
N −1
qh2 (t) = qk φk (t) + qk+ 12 φk+ 12 (t), qh2 (t) ∈ Vh2 ([a, b]). (3.16)
k=0 k=0
In the space Vh2 ([a, b]), the action functional (1.1) becomes
-
b
S (t, qh2 (t)) = L t, qh2 (t), q̇h2 (t) d t
a
−1 - tk+1
N
= L t, qh2 (t), q̇h2 (t) d t
k=0 tk
N −1
= L(tk , qk , qk+ 12 , tk+1 , qk+1 )(tk+1 − tk ), (3.17)
k=0
where
-
1 tk+1
L(tk , qk , qk+ 12 , tk+1 , qk+1 ) = L t, qh2 (t), q̇h2 (t) d t. (3.18)
tk+1 − tk tk
From (3.21) and (3.22), we can solve for qk+ 12 and qk−1+ 12 respectively, then
substitute them into (3.20) and finally solve for qk+1 . Therefore, the discrete Euler–
Lagrange equation (3.20) – (3.22) determines a discrete flow
Ψ : M × M −→ M × M,
Ψ(tk−1 , qk1 , tk , qk ) = (tk , qk , tk+1 , qk+1 ).
ΘLv− (tk , qk , qk+ 12 , tk+1 , qk+1 ) = D2 L(tk , qk , qk+ 12 , tk+1 , qk+1 )(tk+1 − tk )dqk ,
and
ΘLv+ (tk , qk , qk+ 12 , tk+1 , qk+1 ) = D5 L(tk , qk , qk+ 12 , tk+1 , qk+1 )(tk+1 − tk )dqk+1 .
ΘLv− + ΘLv+ = d((tk+1 − tk )L) − D3 L(tk , qk , qk+ 12 , tk+1 , qk+1 )dqk+ 12 . (3.24)
14.3 Discrete Mechanics Based on Finite Element Methods 611
and
-
1
tk+1
G(qk , qk+ 12 , qk+1 ) = V qk fk (t)+qk+1 fk+1 (t)+qk+ 12 fk+ 12 (t) d t,
tk+1 − tk tk
where % &
2(t − tk ) t − tk
fk (t) = −1 −1 ,
hk hk
% &
2(tk+1 − t) tk+1 − t
fk+1 (t) = −1 −1 ,
hk hk
t − tk t − tk
fk+ 12 (t) = 4 1− .
hk hk
For the discrete Lagrangian (3.27), the discrete Euler–Lagrange equations (3.20) –
(3.22) become
a1 qk−1 + a2 qk + a3 qk+1 + a4 qk− 12 + a5 qk+ 12 − d1 hk − d2 hk−1 = 0, (3.28)
8 ∂G(qk , qk+ 12 , qk+1 )
− 2
qk + qk+1 − 2qk+ 12 − = 0, (3.29)
3hk ∂qk+ 1
2
8 ∂G(qk−1 , qk−1+ 12 , qk )
− 2
qk−1 + qk − 2qk−1+ 12 − = 0, (3.30)
3hk−1 ∂qk−1+ 1
2
612 14. Lee-Variational Integrator
where
% &
1 1 7 1 1 1 1
a1 = , a2 = + , a3 = ,
3 hk−1 3 hk−1 hk 3 hk
8 1 8 1 ∂G(qk , qk+ 1 , qk+1 ) ∂G(qk−1 , qk−1+ 1 , qk )
a4 = − , a5 = − , d1 = 2
, d2 = 2
.
3 hk−1 3 hk ∂qk ∂qk
The solution of (3.28) – (3.30) preserves the Lagrange 2-form
* 2
+
1 ∂ G(qk , qk+ 1 , qk+1 )
− hk 2
−M d qk ∧ d qk+1 , (3.31)
3hk ∂ qk ∂ qk+1
where
* +* +
16 ∂ 2 G(qk , qk+ 1 , qk+1 ) 16 ∂ 2 G(qk , qk+ 1 , qk+1 )
2 2
+ hk + hk
3hk ∂qk+ 1 ∂qk 3hk ∂qk+ 1 ∂qk
M= * 2
2 + 2
.
32 ∂ G(qk , qk+ 1 , qk+1 )
− hk 2
2
3hk ∂qk+ 1
2
If we take the fixed time steps hk−1 = hk = h, then (3.28) – (3.30) become
qk−1 − 8qk− 12 + 14qk − 8qk+ 12 + qk+1
− d1 hk − d2 hk−1 = 0, (3.32)
3h2
8 ∂G(qk , qk+ 12 , qk+1 )
− q k + q k+1 − 2q 1 − = 0, (3.33)
3h2 k+ 2 ∂qk+ 12
8 ∂G(qk−1 , qk−1+ 12 , qk )
− 2
qk−1 + qk − 2qk−1+ 12 − = 0, (3.34)
3h ∂qk−1+ 12
which preserve
* +
1 ∂ 2 G(qk , qk+ 12 , qk+1 )
−h −M d qk ∧ d qk+1 , (3.35)
3h ∂qk ∂qk+1
where
* +* +
16 ∂ 2 G(qk , qk+ 1 , qk+1 ) 16 ∂ 2 G(qk , qk+ 1 , qk+1 )
2 2
+h +h
3hk ∂qk+ 1 ∂qk 3hk ∂qk+ 1 ∂qk
M= * 2 2 + 2
.
32 ∂ G(qk , qk+ 1 , qk+1 )
−h 2
2
3h ∂qk+ 1
2
Suppose qk is the solution of (3.28) – (3.30) and q(t) is the solution of (3.10), then
from the convergence theory of finite element methods [Cia78,Fen65] , we have
q(t) − qh2 (t) ≤ Ch3 , (3.36)
where
N
N −1
qh2 (t) = qk φk (t) + qk+ 12 φk+ 12 (t),
k=0 k=0
h = maxk {hk } and C is a constant independent of h.
14.3 Discrete Mechanics Based on Finite Element Methods 613
and the discrete energy evolution equation from the variation δtk
which is a discrete version of (1.23). For a conservative L, (3.39) becomes the discrete
energy conservation law.
From the boundary terms in (3.37), we can define two 1-forms
θL− (wk ) = (D1 L(wk )(tk+1 − tk ) − L(wk ))dtk + D2 L(wk )(tk+1 − tk )dqk , (3.40)
and
θL+ (wk ) = (D3 L(wk )(tk+1 − tk ) + L(wk ))dtk+1 + D4 L(wk )(tk+1 − tk )dqk+1 .
(3.41)
These two 1-forms are the discrete version of the extended Lagrange 1-form (1.29).
Unlike the continuous case, the solution of (3.38) does not satisfy (3.39) in general.
Therefore, we must solve (3.38) and (3.39) simultaneously. Using the same method in
the above section, we can show that the coupled integrator
614 14. Lee-Variational Integrator
For the kind of high order discrete Lagrangian, we can obtain similar formulae.
14.3.4 Conclusions
Recently, it has been proved [GLWW01] that the symplectic structure is preserved not
only on the phase flow but also on the flow with respect to symplectic vector fields
as long as certain cohomological condition is satisfied in both continuous and discrete
cases. This should be able to be extended to the cases in this chapter.
Bibliography
15.1 Introduction
Birkhoffian representation is a generalization of Hamiltonian representation, which
can be applied to hadron physics, statistical mechanics, space mechanics, engineering,
biophysics, etc. Santilli[San83a,San83b] . All conservative or nonconservative, self-adjoint
or non self-adjoint, unconstrained or nonholonomic constrained systems always ad-
mit a Birkhoffian representation (Guo[GLSM01] and Santilli[San83b] ). In last 20 years,
many researchers have studied Birkhoffian mechanics and obtained a series of results
in integral theory, stability of motion, inverse problem, and algebraic and geometric
description, etc.
Birkhoff’s equations are more complex than Hamilton’s equations, and the study
of the computational methods of the former is also more complicated. There are no
result on computational methods for Birkhoffian system before. In general, the known
difference methods are not generally applicable to Birkhoffian system. A difference
scheme used to solve Hamiltonian system should be Hamiltonian scheme (Hairer, Lu-
bich and Wanner[HLW02] and Sanz-Serna and Calvo[SSC94] ), so a difference scheme to
simulate Birkhoffian system should be a Birkhoffian scheme. However, the conven-
tional difference schemes such as Euler center scheme, leap-frog scheme, etc., are not
Birkhoffian schemes. So, a way to systematically construct a Birkhoffian scheme is
necessary, and this is the main context in this chapter.
Both the Birkhoffian and Hamiltonian systems are usually of finite dimensional
(Arnold[Arn89] and Marsden and Ratiu[MR99] ), infinite dimension system has not been
proposed before. The algebraic and geometric profiles of finite dimensional Birkhof-
fian systems are described in local coordinates, and general nonautonomous Hamil-
tonian systems are considered as autonomous Birkhoffian systems (Santilli[San83b] ).
Symplectic schemes are systematically developed for standard Hamiltonian systems
618 15. Structure Preserving Schemes for Birkhoff Systems
and for general Hamiltonian systems on the Poisson manifold, which belong to au-
tonomous and semi-autonomous Birkhoffian systems (Feng and Wang[FW91bandFW91a]
and Feng and Qin[FQ87] ). So, in this chapter, we just discuss the nonautonomous
Birkhoffian system in detail. Thereby, Einstein’s summation convention is used.
In Section 15.2, Birkhoffian systems are sketched out via variational self-adjointness,
with which we shows the relationship between Birkhoffian and Hamiltonian systems
more essentially and directly. Then the basic geometrical properties of Birkhoffian sys-
"
tems are presented. In Section 15.3, the definitions of K(z)-Lagrangian submanifolds
"
is extended to K(z, t)-Lagrangian submanifolds with parameter t. Then the relation-
ship between symplectic mappings and gradient mappings are discussed. In Section
15.4, the generating functions for the phase flow of the Birkhoffian systems are con-
structed and the method to simulate Birkhoffian systems by symplectic schemes of any
order is given. Section 15.5 contains an illustrating example. Schemes of order one,
two, and four are derived for the linear damped oscillator. In the last Section 15.6,
numerical experiments are given.
are further used. Following the terminology suggested by Santilli[San83b] , this is called
Birkhoff’s equation or Birkhoffian system under some additional assumptions. The
function B(z, t) is called the Birkhoffian function because of certain physical differ-
ence with Hamiltonian. Also, the Fi (i = 1, 2, · · · , 2n) are Birkhoffian functions. A
representation of Newton’s equations via Birkhoff’s equation is called a Birkhoffian
representation.
Definition 2.1. Birkhoff’s equations (2.1) are called autonomous when the functions
Fi and B are independent of the time variable. In this case, the equations are of the
simple form
d zj ∂ B(z)
Kij (z) − = 0. (2.2)
dt ∂ zi
They are called semi-autonomous when the functions Fi do not depend explicitly on
time. In this case, the equations have the more general form
d zj ∂ B(z, t)
Kij (z) − = 0.
dt ∂ zi
15.2 Birkhoffian Systems 619
They are called nonautonomous when both the functions Fi and B explicitly depen-
dent on time. Then, the equations read as follow:
d zj ∂ B(z, t) ∂ Fi (z, t)
Kij (z, t) − − = 0. (2.3)
dt ∂ zi ∂t
They are called regular when the functional determinant is unequal to zero in the
region considered, i.e.,
$ ) = 0,
det (Kij )(Re
otherwise, degenerate.
Kij + Kji = 0,
∂ Kij ∂ Kjk ∂ Kki
+ + = 0, (2.5)
∂ zk ∂ zi ∂ zj
∂ Kij ∂ Di ∂ Dj
= − .
∂t ∂ zj ∂ zi
Ω = d (Fi d zi ), (2.6)
this geometric property is fully characterized by the first two equations of the condition
(2.5); i.e., the 2-form (2.6) describes the geometrical structure of the autonomous case
(2.2) of the Birkhoff’s equations, it even sketches out the geometric structure of the
semi-autonomous case.
620 15. Structure Preserving Schemes for Birkhoff Systems
For the case Kij = Kij (z, t), the full set of condition (2.5) must be consid-
ered. The corresponding geometric structure can be better expressed by transition
of the symplectic geometry on the cotangent bundle T ∗ M with local coordinates
zi to the contact geometry on the manifold R × T ∗ M with local coordinates z"i
(i = 0, 1, 2, · · · , 2n), z"0 = t[San83b] . More general formulations of an exact contact
2-form exist, although it is now referred to as a (2n+1)-dimensional space,
2n
=
Ω ij d z"i ∧ d z"j = Ω + 2 Di d zi ∧ d t,
K
i,j=0
where
5 6
0 −DT
=
K , D = (D1 , · · · , D2n )T .
D K
the geometric meaning of the condition of the self-adjointness is then the integrability
condition for the exact contact structure (2.7). Here B can be calculated from
∂B ∂ Fi
− = Di +
∂ zi ∂t
for
∂ ∂ Fi ∂ ∂ Fj
Di + = Dj + .
∂ zj ∂t ∂ zi ∂t
All the above discussion can be expressed via the following property.
Remark 2.3. The functions Fi and B can be calculated according to the rules [AH75]
- 1
1
Fi = zj · Kji (λz, t) d λ,
2 0
- 1
∂ Fi
B= zi · D i + (λz, t) d λ.
0 ∂t
15.3 Generating Functions for K(z, t)-Symplectic Mappings 621
Due to the self-adjointness of Birkhoff’s equations, the phase flow of the system
(2.8) conserves the symplecticity
d d
Ω = (Kij d zi ∧ d zj ) = 0.
dt dt
z,
So denoting the phase flow of the Equation (2.8) with ( t) yields
z,
Kij ( t) d zi ∧ d zj = Kij (z, t) d zi ∧ d zj ,
∂ z T ∂ z
z,
K( t) = K(z, t).
∂z ∂z
In the latter, the algorithm preserving this geometric property of the phase flow in
discrete space will be constructed.
or
" z , z, t, t0 )(Tx L) = 0,
(Tx L)T K(
∂ gT ∂ g
K g(z, t, t0 ), t = K(z, t0 ).
∂z ∂z
A difference scheme approximating the Birkhoffian system (2.8) with step size τ
z k+1 = g k (z k , tk + τ, tk ), k0
The graph of the phase flow of the Birkhoffian system (2.8) is g t (z, t0 ) =
" z , z, t, t0 )-Lagrangian submanifold for
g(z, t, t0 ) which is a K(
gzt (z, t0 )T K g t (z, t0 ), t gzt (z, t0 ) = K(z, t0 ).
Similarly, the graph of the phase flow of standard Hamiltonian system is a J"4n -
Lagrangian submanifold.
Consider the nonlinear transformation with two parameters t and t0 from R4n to
itself,
5 6 5 6 5 6
z
w α1 (
z , z, t, t0 )
α(t, t0 ) : −→ = , (3.1)
z w α2 (
z , z, t, t0 )
5 6 5 6 5 1 6
w z w, t, t0 )
α (w,
−1
α (t, t0 ) : −→ = .
w z α2 (w,
w, t, t0 )
Let α be a diffeomorphism from R4n to itself, then it follows that α carries ev-
"
ery K-Lagrangian submanifold into a J4n -Lagrangian submanifold, if and only if
" i.e.,
α∗T J4n α∗ = K,
5 6T 5 65 6 5 6
Aα Bα J2n O Aα Bα K( z , t) O
= .
Cα Dα O −J2n Cα Dα O −K(z, t0 )
"
Conversely, α−1 carries every J4n -Lagrangian submanifold into a K-Lagrangian sub-
manifold.
Theorem 3.4. Let M ∈ R2n×2n , α given as in (3.1), and define a fractional trans-
formation
under the transversality condition |Cα M + Dα | = 0. Then the following four condi-
tions are mutually equivalent:
|Cα M + Dα | = 0, |M C α − Aα | = 0,
|C α N + Dα | = 0, |N Cα − Aα | = 0.
then there uniquely exists in R" a gradient mapping w → w = f (w, t, t0 ) with Ja-
cobian fw (w, t, t0 ) = N (w, t, t0 ) and a uniquely defined scalar generating function
φ(w, t, t0 ), such that
f (w, t, t0 ) = φw (w, t, t0 ),
α1 (g(z, t, t0 ), z, t, t0 ) = f α2 (g z, t, t0 ), z, t, t0 , t, t0
= φw α2 g(z, t, t0 ), z, t, t0 , t, t0 , (3.3)
and
N = (Aα M + Bα )(Cα M + Dα )−1 ,
M = (Aα N + B α )(C α N + Dα )−1 .
then
∂f ∂ α1 ∂ g ∂ α1 ∂z
N= = + = (Aα M + Bα )(Cα M + Dα )−1 .
∂w ∂ z ∂ z ∂z ∂w
So,
(Aα M + Bα )T (Cα M + Dα ) − (Cα M + Dα )T (Aα M + Bα ) = 0,
Consider the construction of f (w, t, t0 ) and z(w, t, t0 ). Since z(w, t, t0 )◦α2 (g(z, t, t0 ),
z, t, t0 ) ≡ z, substituting w = α2 (g(z, t, t0 ), z, t, t0 ) in (3.4) and (3.5) yields Equa-
tion (3.3).
Theorem 3.6. f (w, t, t0 ) obtained in Theorem 3.5 is also the solution of the following
implicit equation:
α1 f (w, t, t0 ), w, t, t0 = g α2 (f (w, t, t0 ), w, t, t0 ), t, t0 .
15.4 Symplectic Difference Schemes for Birkhoffian Systems 625
Remark 3.8. The proofs of Theorems 3.6 and 3.7 are similar to that of Theorem 3.5
and are omitted here. Similar to Theorem 3.6, the function g(z, t, t0 ) is the solution of
the implicit equation
In Section 15.2, it is indicated that for a general Birkhoffian system, there exists the
common property that its phase flow is symplectic. With the result in the last sec-
tion, symplectic schemes for Birkhoffian systems are constructed by approximating
the generating functions.
Birkhoff’s phase flow is denoted by g t (z, t0 ) and it is a one-parameter group of
K(z, t)-symplectic mappings at least local in z and t, i.e., g t0 = identity, g t1 +t2 =
g t1 ◦ g t2 . Here z is taken as an initial value when t = t0 , and z(z, t, t0 ) = g t (z, t0 ) =
g(t; z, t0 ) is the solution of the Birkhoffian system (2.8).
Theorem 4.1. Let α be defined as in Theorem 3.5. Let z → z = g t (z, t0 ) be the phase
flow of the Birkhoffian system (2.8), M (t; z, t0 ) = gz (t; z, t0 ) is its Jacobian. At some
initial point z, i.e., t = t0 , z = z, if
then for sufficiently small |t − t0 | and in some neighborhood of z ∈ R2n there exists
a gradient mapping w → w = f (w, t, t0 ) with symmetric Jacobian fw (w, t, t0 ) =
N (w, t, t0 ) and a uniquely determined scalar generating function φ(w, t, t0 ) such that
626 15. Structure Preserving Schemes for Birkhoff Systems
so
∂w
dw ∂w dw
∂w ∂F ∂ α1 ∂ α2
∂w
= − = Aα − Cα K −1 ∇ B + + − .
∂t dt ∂ t dt ∂t ∂t ∂t ∂w ∂t
∂w
Since = 0, so w = w(w,
t) exists and is solvable in (w),
but it cannot be solved
∂w
explicitly from the transformation α and α−1 , we have
∂w
∂w
Ā z, z, , t, t0 = ,
∂w ∂t
and the Equations (4.4) and (4.5). Then, from (4.6), the Equation (4.3) follows.
15.4 Symplectic Difference Schemes for Birkhoffian Systems 627
Remark 4.2. Because of the forcing term in (2.1), the Hamilton–Jacobi equation
for the generating function φ(w, t, t0 ) cannot directly be derived, but instead the
Hamilton–Jacobi equation (4.3) for φw (w, t, t0 ) can be easily derived. Assume the
generating function φw (w, t, t0 ) can be expanded as a convergent power series in t
∞
φw (w, t, t0 ) = (t − t0 )k φ(k)
w (w, t0 ). (4.8)
k=0
Lemma 4.3. The k-th order total derivative of A defined as in Theorem 4.1 with re-
spect to t can be described as
∞
∞
Dtk A = ∂φw A (t − t0 )i φ(k+i)
w + ∂ φww A (t − t0 )i φ(k+i)
ww
i=0 i=0
∞
∞
+∂t ∂φw A (t − t0 )
i
φ(k−1+i)
w + ∂t ∂φww A (t − t0 )i φ(k−1+i)
ww
i=0 i=0
k
k−m
k−m−n
+ Cm
k Cnk−m ∂φnw ∂φl ww ∂tm A
m=0 n=1 l=1 h1 +···+hn
+j1 +···+jl =k−m
∞ ∞
(h1 +i) (hn +i)
· (t − t0 )i φw ,· · ·, (t − t0 )i φw ,
i=0 i=0
∞
∞
· (t − t0 )i φ(j
ww
1 +i)
, · · ·, (jl +i)
(t − t0 )i φw ,
i=0 i=0
(k) (k)
Dtk At0 = ∂φw At0 φw + ∂φww At0 φww
(k−1) (k−1)
+∂t ∂φw At0 φw + ∂t ∂φw w At0 φww
k
k−m
k−m−n
+ Cm
k Cnk−m ∂φnw ∂φl ww ∂tm At0
m=0 n=1 l=1 h1 +···+hn
+j1 +···+jl =k−m
(h ) (h ) (j1 ) (jl )
· φw 1 , · · · , φw n , φww , · · · , φww ,
(0) (0)
where At0 = A(φw , w, φww , t0 , t0 ).
By means of the representations of the total derivative of A, the following results
are proved.
Theorem 4.4. Let A and α be analytic. Then the generating function φwα,A (w, t, t0 )
= φw (w, t, t0 ) can be expanded as a convergent power series in t for sufficiently small
|t − t0 |
∞
φw (w, t, t0 ) = (t − t0 )k φ(k)
w (w, t0 ), (4.9)
k=0
(k)
and φw (k ≥ 0), can be recursively determined by the following equations
φ(0)
w (w, t0 ) = f (w, t0 , t0 ), (4.10)
(0)
φ(1) (0)
w (w, t0 ) = A φw , w, φww , t0 , t0 , (4.11)
1
φk+1
w (w, t0 ) = Dtk A φ(0) (0)
w , w, φww , t0 , t0 . (4.12)
(k + 1) !
By Equation (4.2),
Using Equation (4.3) and comparing (4.15) with (4.14), we get (4.11) and (4.12).
15.5 Example 629
Theorem 4.5. Let A and α be analytic. For sufficiently small time-step τ > 0, take
m
(m)
ψw (w, t0 + τ, t0 ) = τ i φ(i)
w (w, t0 ), m = 1, 2, · · · ,
i=0
(i)
where φw are determined by Equations (4.10) – (4.12).
(m)
Then, ψw (w, t0 + τ, t0 ) defines a K(z,t)-symplectic difference scheme z = z k →
z k+1 = z,
(m)
α1 (z k+1 , z k , tk+1 , tk ) = ψw α2 (z k+1 , z k , tk+1 , tk ), tk+1 , tk (4.16)
|C α N (m) (w, t0 + τ, t0 ) + Dα | = 0,
where
15.5 Example
In this section, an example illustrates how to obtain schemes preserving the symplectic
structure for a nonconservative system expressed in Birkhoffian representation. Con-
sider the linear damped oscillator
r̈ + ν ṙ + r = 0. (5.1)
5 65 6 5 6
0 −eνt ṙ νeνt p + eνt r
= . (5.2)
eνt 0 ṗ eνt p
and ⎡ ⎤
1
0 1 0
⎢ 2 ⎥
⎢ 1 −νt ⎥
⎢ e 0 0 −e−νt ⎥
⎢ ⎥
α∗−1 =⎢ 2
⎥.
⎢ −
1 ⎥
⎢ 0 1 0 ⎥
⎣ 2 ⎦
1
− e−νt0 0 0 −e−νt0
2
Consequently, using (5.5), (5.6) and (5.2), we derive
⎡ ⎤
5 6 5 6 1
νe νt
p + e νt ˙
p −e νt
q − eνt P − eνt Q
dw ⎢ 2 ⎥
= = =⎣ ⎦,
dt ˙q p 1 − e−νt P
e−νt Q
2
⎡ ⎤
1 −νt 1 −νt
e Q− e P
dw ⎢ 4 2 ⎥
=⎣ ⎦.
dt 1 νt 1
e P + eνt Q
4 2
(0) (1)
= φw + φw τ , so the first order scheme for the system (5.2) reads as follows:
Set w
= eνt/2 p − eνt0 /2 p,
Q P = −eνt/2 q + eνt0 /2 q,
1 1
Q = (eνt/2 q + eνt0 /2 q), P = − (eνt p + eνt0 p).
2 2
The Jacobian of α is
⎡ ⎤
0 eνt/2 0 −eνt0 /2
⎢ −eνt/2 eνt0 /2 ⎥
⎢ 0 0 ⎥
⎢ ⎥
α∗ = ⎢ 1 νt/2 1 νt0 /2 ⎥
⎢ e 0 e 0 ⎥
⎣ 2 2 ⎦
1 1
0 − eνt 0 − eνt0
2 2
and the inverse ⎡ ⎤
1
0 1 0
⎢ 2 ⎥
⎢ ⎥
⎢ 1 −νt
e 0 0 −e −νt ⎥
⎢ ⎥
=⎢ ⎥.
2
α∗−1 ⎢ ⎥
⎢ 1 ⎥
⎢ 0 − 1 0 ⎥
⎣ 2 ⎦
1
− e−νt0 0 0 −e−νt0
2
Direct calculation yields the scheme of second order
e−ντ /2
Abbreviating the matrix (∗) in (5.8) by M (τ ), then by composition[Yos90,QZ92]
Δ
we have the scheme of order four
5 6 5 6
qk+1 qk
= M (c1 τ )M (c2 τ )M (c1 τ ) , (5.9)
pk+1 pk
where
1 −21/3
c1 = , c2 = .
2 − 21/3 2 − 21/3
If take m = 2, we have
φ(2)
w = 0.
15.5 Example 633
Now take m = 3,
1 ∂ ∂ dw " ∂ " ∂w
dw " " ∂
∂w "
dw ∂ w
φ(3)
w = + −
3! ∂ t ∂ t dt ∂w " dt ∂ t ∂w∂w " dt ∂ t
∂w" ∂ dw ∂ ∂w " dw
− − . (5.10)
∂ w ∂ t dt ∂ t ∂ w dt
For equation q̈ + ν q̇ + q = 0, 3rd derivatives of φ in time t = t0 , only one term to
appear, i.e.,
∂ ∂w" ∂ "
dw ∂ w
− .
∂t ∂w ∂w" dt ∂ t
Simple calculation yields
1 ν
−1 −
ν− −ν P − Q
(3) 1 8 8
φw 2
2
t=t0
= − ν ν 1 ν
6 − − −1 − Q−P
2 8 4 2
1 1
− + ν2 Q +
νP
1 4 16 2
= −
6 1 1 2 νQ
− + ν +P
4 16 2
2 2
ν ν ν
− 2 Q + − 2 P
2 2 2
= 2
,
ν2 ν ν
−2 P + −2 Q
2 2 2
(1) (3)
" = φw Δt + φw Δt2 , i.e.,
we get 4-th order symmetrical symplectic scheme: w
eνtk+1 /2 qk+1 − eνtk /2 qk eνtk+1 /2 pk+1 + eνtk /2 pk eνtk+1 /2 qk+1 − eνtk /2 qk
= +ν
τ 2 4
1 ν2
+τ 2 − 2 eνtk+1 /2 pk+1 + eνtk /2 pk
24 × 4 2
2
ν ν νtk+1 /2
+ −2 e qk+1 + eνtk /2 qk ,
2 2
eνtk+1 /2 pk+1 − eνtk /2 pk eνtk+1 /2 qk+1 + eνtk /2 qk eνtk+1 /2 pk+1 + eνtk /2 pk
= −ν
τ 2 4
1 ν2
−τ 2 − 2 eνtk+1 /2 qk+1 + eνtk /2 qk
24 × 4 2
ν2 ν νtk+1 /2
+ −2 e pk+1 + eνtk /2 pk .
2 2
(5.11)
This method is easily extended to more general ODEs such as
ṗ + β (t)p + V (q, t) = 0,
(5.12)
q̇ − G(p, t) = 0.
634 15. Structure Preserving Schemes for Birkhoff Systems
Remark 5.1. The derived schemes (5.7), (5.8), and (5.9) are K(z, t)-symplectic, i.e.,
for τ > 0 and k ≤ 0 they satisfy the Birkhoffian condition
In this section, we present numerical results for the linear damped oscillator (5.1),
resp., (5.2) using the derived K(z, t)-symplectic schemes (5.7), (5.8), and (5.9) of
order one, two, and four, respectively. Further, we use Euler’s midpoint scheme (5.4),
which is not K(z, t)-symplectic but shows convenient numerical results[MW01] , and
further Euler’s explicit scheme for comparison.
In the presented figures, the initial values are always chosen as q(0) = 1,
p(0) = q̇(0) = −1, and the time interval is from 0 to 25. There are only small
differences in the behavior of the different schemes choosing other initial values. The
actual error, err = |approximate solution - true solution|, is computed with step size
τ = 0.2. Using different step sizes, the schemes always show the same quality, which
is emphasized by representing the results in a double logarithmic scale using step sizes
τ = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5. The orbits are computed with step size τ = 0.05.
The first comparison is given between scheme (5.7) and Euler’s explicit scheme
both are of order one. For smaller ν, i.e., 0 ≤ ν ≤ 1.3 scheme (5.7) is better, and for
ν > 1.3 Euler’s explicit scheme is better. The second comparison is given between
scheme (5.8) and Euler’s midpoint scheme (5.4) both are of order two. For 0 ≤ ν ≤
0.5 both schemes show the same behavior, for 0.5 < ν < 2.8 scheme (5.8) is better,
where the most advantage is around ν = 2, and for 2.8 ≤ ν Euler’s midpoint scheme
behaves better. The third comparison is given between scheme (5.9) of order four and
scheme (5.8) of order two. Both schemes have the same structure preserving property,
and therefore the higher order scheme (5.9) shows a clear superiority over the two-
order scheme. These differences between the discussed schemes are illustrated by the
error curves (Figs. 6.1 and 6.4).
For the energy function (5.3), the comparisons of the energy error H, between the
different schemes are also done in double logarithmic scales (Figs. 6.5 and 6.8). The
result shows that the dominance is not clear between scheme (5.7) and Euler’s explicit
scheme while scheme (5.8) is always better than Euler’s midpoint scheme for growing
ν, even for ν ≥ 2.8. Scheme (5.9) keeps its superiority in the comparisons.
The comparisons also show that it is possible for different schemes obtained from
different transformation α, that different quantities are preserved. This point is proved
to be true in the generating function method for Hamiltonian systems (see Feng et
al[FW91b,FW91a] ). The extension to application in Birkhoffian systems will also be stud-
ied in a prospective paper.
15.6 Numerical Experiments 635
0
q’’ + 0.6 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
scheme (5.7)
−2
scheme (5.8)
10 scheme (5.9)
−4
10
max−err
−6
10
−8
10
−10
10 −2 −1 0
10 10 10
tau
Fig. 6.1. Error comparison between the different schemes for ν = 0.6
0
q’’ + 1.3 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
scheme (5.7)
−2
scheme (5.8)
10 scheme (5.9)
−4
10
max−err
−6
10
−8
10
−10
10 −2 −1 0
10 10 10
tau
Fig. 6.2. Error comparison between the different schemes for ν = 1.3
636 15. Structure Preserving Schemes for Birkhoff Systems
0
q’’ + 1.9 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
−2
10
scheme (5.7)
scheme (5.8)
scheme (5.9)
−4
10
−6
10
max−err
−8
10
−10
10
−12
10
−14
10 −2 −1 0
10 10 10
tau
Fig. 6.3. Error comparison between the different schemes for ν = 1.9
0
q’’ + 2.8 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
scheme (5.7)
−2
scheme (5.8)
10 scheme (5.9)
−4
10
max−err
−6
10
−8
10
−10
10 −2 −1 0
10 10 10
tau
Fig. 6.4. Error comparison between the different schemes for ν = 2.8
15.6 Numerical Experiments 637
0
q’’ + 2.8 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
scheme (5.7)
−2
scheme (5.8)
10 scheme (5.9)
−4
10
max−err−H
−6
10
−8
10
−10
10 −2 −1 0
10 10 10
tau
Fig. 6.5. Energy error comparison between the different schemes for ν = 0.6
0
q’’ + 1.3 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
scheme (5.7)
−2
10 scheme (5.8)
scheme (5.9)
−4
10
max−err−H
−6
10
−8
10
−10
10
−12
10 −2 −1 0
10 10 10
tau
Fig. 6.6. Energy error comparison between the different schemes for ν = 1.3
638 15. Structure Preserving Schemes for Birkhoff Systems
0
q’’ + 1.9 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
−2
10
scheme (5.7)
scheme (5.8)
scheme (5.9)
−4
10
max−err−H
−6
10
−8
10
−10
10
−12
10
−14
10 −2 −1 0
10 10 10
tau
Fig. 6.7. Energy error comparison between the different schemes for ν = 1.9
0
q’’ + 2.8 * q’ + q = 0 ( tau = 0.01, 0.02, 0.05, 0.1, 0.2, 0.5 )
10
expl Euler
midpoint (5.4)
scheme (5.7)
−2
scheme (5.8)
10 scheme (5.9)
−4
10
max−err−H
−6
10
−8
10
−10
10 −2 −1 0
10 10 10
tau
Fig. 6.8. Energy error comparison between the different schemes for ν = 2.8
Bibliography
[AH75] R.W. Atherton and G.M. Homsy: On the existence and formulation of variational prin-
ciples for nonlinear differential equations. Studies in Applied Mathematics, LIV(1):1531–
1551, (1975).
[Arn89] V. I. Arnold: Mathematical Methods of Classical Mechanics. Springer-Verlag, GTM
60, Berlin Heidelberg, Second edition, (1989).
[FQ87] K. Feng and M.Z. Qin: The symplectic methods for the computation of Hamiltonian
equations. In Y. L. Zhu and B. Y. Guo, editors, Numerical Methods for Partial Differential
Equations, Lecture Notes in Mathematics 1297, pages 1–37. Springer, Berlin, (1987).
[FW91a] K. Feng and D.L. Wang: A Note on conservation laws of symplectic difference
schemes for Hamiltonian systems. J. Comput. Math., 9(3):229–237, (1991).
[FW91b] K. Feng and D.L. Wang: Symplectic difference schemes for Hamiltonian systems in
general symplectic structure. J. Comput. Math., 9(1):86–96, (1991).
[GLSM01] Y.X. Guo, S.K. Luo, M. Shang, and F.X. Mei: Birkhoffian formulations of non-
holonomic constrained systems. Reports on Mathematical Physics, 47:313–322, (2001).
[HLW02] E. Hairer, Ch. Lubich, and G. Wanner: Geometric Numerical Integration. Num-
ber 31 in Springer Series in Computational Mathematics. Springer-Verlag, Berlin, (2002).
[MP91] E. Massa and E. Pagani: Classical dynamics of non-holonomic systems : a geomet-
ric approach. Annales de l’institut Henri Poincar (A) Physique thorique, 55(1):511–544,
(1991).
[MR99] J. E. Marsden and T. S. Ratiu: Introduction to Mechanics and Symmetry. Number 17
in Texts in Applied Mathematics. Springer-Verlag, Berlin, second edition, (1999).
[MW01] J. E. Marsden and M. West: Discrete mechanics and variational integrators. Acta
Numerica, 10:357–514, (2001).
[QZ92] M. Z. Qin and W. J. Zhu: Construction of higher order symplectic schemes by com-
position. Computing, 47:309–321, (1992).
[San83a] R.M. Santilli; Foundations of Theoretical Mechanics I. Springer-Verlag, New York,
Second edition, (1983).
[San83b] R.M. Santilli: Foundations of Theoretical Mechanics II. Springer-Verlag, New York,
Second edition, (1983).
[SQ03] H. L. Su and M. Z. Qin: Symplectic schemes for Birkhoffian system. Technical Report
arXiv: math-ph/0301001, (2003).
[SSC94] J. M. Sanz-Serna and M. P. Calvo: Numerical Hamiltonian Problems. AMMC 7.
Chapman & Hall, London, (1994).
[SSQS07] H. L. Su, Y.J. Sun, M. Z. Qin, and R. Scherer: Symplectic schemes for Birkhoffian
system. Inter J of Pure and Applied Math, 40(3):341–366, (2007).
[SVC95] W. Sarlet, A. Vandecasteele, and F. Cantrijn: Derivations of forms along a map:
The framework for time-dependent second-order equations. Diff. Geom. Appl., 5:171–203,
(1995).
[Yos90] H. Yoshida: Construction of higher order symplectic integrators. Physics Letters A,
150:262–268, (1990).
Chapter 16.
Multisymplectic and Variational Integrators
16.1 Introduction
The introduction of symplectic integrators is a milestone in the development of numer-
ical analysis[Fen85] . It has led to the establishment of structure-preserving algorithms, a
very promising subject. Due to its high accuracy, good stability and, in particular, the
capability for long-term computation, the structure-preserving algorithms have proved
to be very powerful in numerical simulations. The applications of structure-preserving
algorithms can be found on diverse branches of physics, such as celestial mechanics,
quantum mechanics, fluid dynamics, geophysics[LQHD07,MPSM01,WHT96] , etc.
Symplectic algorithms for finite dimensional Hamiltonian systems have been well
established. They not only bring new insights into existing methods but also lead to
many powerful new numerical methods. The structure-preserving algorithms for in-
finite dimensional Hamiltonian systems are comparatively less explored. Symplec-
tic integrators for infinite dimensional Hamiltonian systems were also considered
[Qin90,LQ88,Qin87,Qin97a]
. The basic idea is, first to discretize the space variables appro-
priately so that the resulting semi-discrete system is a Hamiltonian system in time;
and second, to apply symplectic methods to this semi-discrete system. The symplectic
integrator obtained in this way preserves a symplectic form which is a sum over the
discrete space variables. In spite of its success, a problem remains: the change of the
symplectic structure over the spatial domain is not reflected in such methods.
This problem was solved by introducing the concept of multisymplectic integra-
tors (Bridges and Reich[BR01a,BR06] ). In general, an infinite dimensional Hamiltonian
system can be reformulated as a multisymplectic Hamiltonian system in which as-
sociated to every time and space direction, there exists a symplectic structure and a
642 16. Multisymplectic and Variational Integrators
1. Multisymplectic geometry Exclusively, local coordinates are used and the notion
of prolongation spaces instead of jet bundles[Olv86,Che05c] is employed. The covariant
configuration space is denoted by X × U and X represents the space of independent
variables with coordinates xμ (μ = 1, 2, · · · , n, 0), and U the space of dependent
variables with coordinates uA (A = 1, 2, · · · , N ). The first-order prolongation of X ×
U is defined to be
U (1) = X × U × U1 , (2.1)
where U1 represents the space consisting of first-order partial derivatives of uA with
respect to xμ .
Let φ : X → U be a smooth function, then its first prolongation is denoted by
pr1 φ = (xμ , φA , φA
μ ).
where
-
∂L ∂L ∂L
A = + Dt φt − L + Dx φt ξ0
M ∂t ∂ φt ∂ φx
∂L ∂L ∂L
+ + Dx φx − L + Dt φx ξ1
∂x ∂ φx ∂ φt
!
∂L ∂L ∂L
+ − Dx − Dt α d x ∧ d t, (2.4)
∂φ ∂ φx ∂ φt
644 16. Multisymplectic and Variational Integrators
and
-
∂L ∂L
B = φt − L d x − φt d t ξ 0
∂M ∂ φt ∂ φx
∂L ∂L
+ L− φx d t + φx d x ξ 1
∂ φx ∂ φt
!
∂L ∂L
+ dt − dx α . (2.5)
∂ φx ∂ φt
If ξ 1 (x), ξ 0 (x), and α(x, t, φ(x, t)) have compact support on M , then B = 0. In this
case, with the requirement of δS = 0 and from (2.4), the variation ξ 0 yields the local
energy evolution equation
∂L ∂L ∂L
+ Dt φt − L + Dx φt = 0, (2.6)
∂t ∂ φt ∂ φx
For a conservative L, i.e., the one that does not depend on x, t explicitly, (2.6) and
(2.7) become the local energy conservation law and the local momentum conservation
law respectively.
The variation α yields the Euler–Lagrange equation
∂L ∂L ∂L
− Dx − Dt = 0. (2.8)
∂φ ∂φx ∂φt
If the condition that ξ 1 (x, t), ξ 0 (x, t), α(x, t, φ(x, t)) have compact support on M is
not imposed, then from the boundary integral B, we can define the Cartan form
∂L
∂L ∂L ∂L
ΘL = dφ ∧ dt − dφ ∧ dx + L − φx − φt d x ∧ d t, (2.9)
∂φx ∂ φt ∂ φx ∂ φt
which satisfies using the interior product and the pull-back mapping ()∗ ,
-
1 ∗ 1
B= pr φ pr V ΘL . (2.10)
∂M
Theorem 2.1. [MPS98,GAR73] Suppose φ is the solution of (2.8), and let η λ and ζ λ be
two one-parameter symmetry groups of Equation (2.8), and V1 and V2 be the corre-
sponding infinitesimal symmetries, then we have the multisymplectic form formula
-
1 ∗ 1
pr φ pr V1 pr1 V2 ΩL = 0. (2.11)
∂M
16.2 Multisymplectic Geometry and Multisymplectic Hamiltonian Systems 645
M z t + Kz x = (z S(z), (2.12)
1 1
ω= d z ∧ M d z, κ= d z ∧ K d z,
2 2
which are associated to the time direction and the space direction, respectively.
The system (2.12) satisfies a local energy conservation law
Dt E + Dx F = 0, (2.14)
Dt I + Dx G = 0 (2.15)
∂L ∂L
L = L(φ, φx , φt ) ⇐⇒ H = L − φx − φt ,
∂ φx ∂ φt
∂L ∂L ∂L
− Dx − Dt = 0 ⇐⇒ M z t + Kz x = (z S(z),
∂φ ∂ φx ∂ φt
-
(pr1 φ)∗ (pr1 V1 pr1 V2 ΩL ) = 0 ⇐⇒ Dt ω + Dx κ = 0,
∂M
% & % &
∂L ∂L
Dt φt − L + Dx φt = 0 ⇐⇒ Dt E + Dx F = 0,
∂ φt ∂ φx
% & % &
∂L ∂L
Dx φx − L + Dt φx = 0 ⇐⇒ Dt I + Dx G = 0.
∂ φx ∂ φt
z j+1 − z ji+ 1 j+ 1
z 2 − zi
j+ 1
2
i+ 1 j+ 1
M 2 2
+ K i+1 = ∇z S z i+ 12 , (3.1)
Δt Δx 2
where Δ t and Δ x are the time step size and space step size, respectively, and
1
z ji ≈ z(iΔz, jΔt), z j+1 1 = z j+1
+ z j+1
,
i+ 2 2 i
i+1
j+ 12 1 j j j+1 j+1
z i+ 1 = z i + z i+1 + z i + z i+1 , etc.
2 4
M1 z t + K1 z x = ∇z S1 (z), (3.5)
where
1
z = (u, v, w)T , S1 (z) = (v 2 − w2 ) − cos (u)
2
and ⎛ ⎞ ⎛ ⎞
0 −1 0 0 0 1
⎜ ⎟ ⎜ ⎟
M1 = ⎝ 1 0 0 ⎠, K1 = ⎝ 0 0 0 ⎠.
0 0 0 −1 0 0
Applying the multisymplectic integrator (3.1) to (3.3) yields
j+1 j j+ 1 j+ 1
vi+ 1 − vi+ 1 w 2 − wi 2
j+ 1
− 2 2
+ i+1 = sin ui+ 12 ,
Δt Δx 2
uj+1
i+ 1
− uji+ 1 j+ 1
2 2
= vi+ 12 , (3.6)
Δt 2
j+ 1 j+ 1
u 2 − ui 2
j+ 1
− i+1 = −wi+ 12 .
Δx 2
1 j−1 1 j−1 j
ūj−1
i−1 = u + uj−1 + uji + uji−1 , ūj−1 = u + uj−1 j
i+1 + ui+1 + ui .
4 i−1 i i 4 i
M2 zt + K2 zx = (z S2 (z), (3.9)
where
1 2
z = (p, q, v, w)T , S2 (z) = v + w2 + V (p2 + q 2 )
2
and ⎛ ⎞ ⎛ ⎞
0 1 0 0 0 0 −1 0
⎜ ⎟ ⎜ ⎟
⎜ −1 0 0 0 ⎟ ⎜ 0 0 0 −1 ⎟
M2 = ⎜ ⎟
⎜ 0 0 0 0 ⎟, K2 = ⎜
⎜ 1 0
⎟.
⎝ ⎠ ⎝ 0 0 ⎟⎠
0 0 0 0 0 1 0 0
From the multisymplectic Preissman integrator (3.1), we obtain a six-point integrator
for (3.8)
j+1 j j+ 1 j+ 1 j+ 1
ψ[i] − ψ[i] ψi+12 − 2ψi 2 + ψi−12 1
i + 2
+ Gi,j = 0, (3.10)
Δt Δx 2
where
r 1
ψ[i] = ψi−1,r + 2ψi,r + ψi+1,r , r = j, j + 1,
4
j+ 1 2 j+ 1 j+ 1 2 j+ 1
Gi,j = V ψi− 12 ψi− 12 + V ψi+ 12 ψi+ 12 .
2 2 2 2
M3 z t + K3 z x = (z S3 (z), (3.12)
where
1 2
z = (φ, u, v, w)T , S3 (z) = v + u2 − uw
2
and ⎛ ⎞
1 ⎛ ⎞
0 0 0 0 0 0 1
⎜ 2 ⎟
⎜ 1 ⎟ ⎜ ⎟
⎜ − 0 0 0 ⎟ ⎜ 0 0 −1 0 ⎟
M3 = ⎜
⎜ 2
⎟,
⎟ K3 = ⎜
⎜ 0 1
⎟.
⎜ 0 0 ⎟
0 0 ⎟ ⎝ ⎠
0
⎝ 0 ⎠
−1 0 0 0
0 0 0 0
From the multisymplectic Preissman integrator (3.1), we obtain an eight-point inte-
grator
16.3 Multisymplectic Integrators and Composition Methods 649
j+ 1 j+ 1 j+ 1 j+ 1
uj+1 j
(i) − u(i) ū2 − ū2i−1 u 2 − 3ui 2 + 3ui−12 − ui−22
+ 3 i+1 + i+1 = 0, (3.13)
Δt 2Δ x Δ x3
where
1
i ψt + ψxx + ψϕ = 0,
2
√ (3.14)
ϕtt − ϕxx + ϕ − |ψ|2 = 0, i= −1
where ψ0 (x), ϕ0 (x) and ϕ1 (x) are known functions. The problems (3.14), (3.15) and
(3.16) has conservative quantity
- xR
2
ψ = ψ ψ̄ d x = 1.
xL
Setting ψ = p + i q, ψx = px + i qx = f + i g,
pt = v, ϕx = w, z = (p, q, f, ϕ, v, w)T .
⎧ 1
⎪
⎪ qt + fx = −ϕp,
⎪
⎪ 2
⎪
⎪
⎪
⎪ 1
⎪
⎪ pt + gx = −ϕq,
⎪
⎪ 2
⎪
⎪
⎪
⎪ 1 1
⎪
⎪ − px = f,
⎪
⎪ 2 2
⎪
⎨
1 1
− qx = − g, (3.17)
⎪
⎪ 2 2
⎪
⎪
⎪
⎪ 1 1 1 1
⎪
⎪ − vt + wx = ϕ − (p2 + q 2 ),
⎪
⎪ 2 2 2 2
⎪
⎪
⎪
⎪ 1 1
⎪
⎪ ϕt = v,
⎪
⎪ 2 2
⎪
⎪
⎪ 1
⎩ 1
− ϕx = − w.
2 2
System (3.17) can be written in standard Bridge form
∂z ∂z
M +K = ∇ S, (3.18)
∂t ∂x
where matrices M and K (3.18) are
⎛ ⎞ ⎛ ⎞
0 −2 0 0 0 0 0 0 0 1 0 0 0 0
⎜ ⎟ ⎜ 0 ⎟
⎜ 2 0 0 0 0 0 0 ⎟ ⎜ 0 0 0 1 0 0 ⎟
⎜ ⎟ ⎜ ⎟
⎜ 0 0 0 0 0 0 0 ⎟ ⎜ −1 0 0 0 0 0 0 ⎟
⎜ ⎟ ⎜ ⎟
1⎜ ⎟ 1⎜ ⎟
M= ⎜ 0 0 0 0 0 0 0 ⎟, K= ⎜ 0 −1 0 0 0 0 0 ⎟
2⎜ ⎟ 2⎜ ⎟
⎜ 0 0 0 0 0 −1 0 ⎟ ⎜ 0 0 0 0 0 0 1 ⎟
⎜ ⎟ ⎜ ⎟
⎜ ⎟ ⎜ ⎟
⎝ 0 0 0 0 1 0 0 ⎠ ⎝ 0 0 0 0 0 0 0 ⎠
0 0 0 0 0 0 0 0 0 0 0 −1 0 0
ω(z) = −2 d p ∧ d q − d ϕ ∧ d v,
κ(z) = d p ∧ d f + d q ∧ d q + d ϕ ∧ d w,
1 1
E(z) = − ϕ(p2 + q 2 ) + (ϕ2 + v 2 − pfx − qgx − ϕwx ),
2 4
1 (3.19)
F (z) = (pft + qgt + ϕwt − f pt − gqt − rw),
4
1 1 1
I(z) = − ϕ(p2 + q 2 ) + (ϕ2 − w2 − f 2 − g 2 + ϕvt ) + (pqt − qpt ),
2 4 2
1
G(z) = (−2pg + 2qf − qvx + vw).
4
16.3 Multisymplectic Integrators and Composition Methods 651
Recently, many math physical equations can be solved by Multisymlectic methods, such
as Gross–Pitaevskii equation[TM03,TMZM08] , Maxwell’s equations[SQS07,SQ03,CYWB06,STMM07] ,
Camassa–Holm equation[Dah07] , Kadomtsev–Petviashvili equation[JYJ06] , Seismic
wave equation[Che04b,Che04a,Che07a,Che07c,Che07b] , Dirac equation[HL04] , and nonlinear
“good” Boussinesq equation[HZQ03,Che05a] , etc.
Now, let us discuss the composition method for constructing high order multisym-
plectic integrators[Che05c,CQ03] . First, recall the definition of a composition method for
ODEs[Yos90,QZ92,Suz92] : Suppose there are n integrators with corresponding operators
s1 (τ ), s2 (τ ), · · ·, sn (τ ) of corresponding order p1 , p2 , · · · , pn , respectively, having
maximal order μ =maxi (pi ). If there exists constants c1 , c2 , · · · , cn such that the or-
der of the integrator whose operator is the composition s1 (c1 τ )s2 (c2 τ ) · · · sn (cn τ )
is m > μ, then the new integrator is called composition integrator of the original n
integrators. This construction of higher order integrators from the lower order ones is
called the composition method.
While constructing higher order integrators, the main task is to determine con-
stants c1 , c2 , · · ·, cn such that the scheme with the corresponding operator
has order m > μ. Now, we will present the basic formula for determining the constants
ci (i = 1, · · · , n). For this purpose, we introduce the symmetrization operator S
p!q!
S(xp z q ) = Pm (xp z q ),
(p + q) !
Pm
n
P S(xn1 1 xn2 2 xn3 3 · · ·) = 0, ci = 1, (3.20)
i=1
where Δx and Δt are the grid sizes in direction x and t, and M is a subset of X. In
this chapter, only an equally spaced grid is considered.
Now for brevity of notations, let M = [a, b] × [c, d] be a rectangular domain and
consider a uniform rectangular subdivision
a = x0 < x1 < · · · < xM −1 < xM = b, c = t0 < t1 < · · · < tN −1 < tN = d,
xi = a + i Δ x, tj = c + j Δ t, i = 0, 1, · · · , M, j = 0, 1, · · · , N,
M Δ x = b − a, N Δ t = d − c.
(4.4)
16.4 Variational Integrators 653
For autonomous Lagrangian and uniform rectangular subdivisions, the discrete action
functional takes the form
−1 N
−1
M
S(ϕ) = L ϕij , ϕi+1j , ϕi+1j+1 , ϕij+1 Δ x Δ t. (4.5)
i=0 j=0
Using the discrete variational principle, we obtain the discrete Euler–Lagrange equa-
tion (variational integrator)
etc. For the discrete Lagrangian, the discrete Euler–Lagrange equation (4.6) is a nine-
point variational integrator. The following results demonstrate the equivalence of vari-
ational integrators and multisymplectic integrators. Consider the sine-Gordon equa-
tion (3.3), then the Lagrangian is given by
1 1
L(u, ux , ut ) = u2x − u2t − cos (u). (4.9)
2 2
The discrete Euler–Lagrange equation (4.6) corresponding to (4.9) is just the nine-
point integrator (3.7). Consider the nonlinear Schrödinger equation (3.8), then the
Lagrangian for (3.8) is given by
1@ 2 A
L(p, q, px , qx , pt , qt ) = px + qx2 + pqt − qpt − V (p2 + q 2 ) . (4.10)
2
j+1 j−1 j+ 1 j− 1 j+ 1 j− 1 j+ 1 j− 1
ψ[i] − ψ[i] ψi+12 + ψi+12 − 2ψi 2
− 2ψi 2
+ ψi−12 + ψi−12
i +
2Δ t Δ x2
1 1
+ Gi,j + Gi,j−1 = 0. (4.11)
4 4
The integrator (4.11) is equivalent to the integrator (3.10), since replacing j by j − 1
in (3.10) and adding the resulting equation to (3.10) leads to (4.11) (see [CQ03] ).
it follows
d qj
− (D1 v)j = 2(p2j + qj2 )pj ,
dt
dp
− j − (D1 w)j = 2(p2j + qj2 )qj , (5.1)
dt
(D1 p)j = vj ,
(D1 q)j = wj ,
N −1
d
ωj + (D1 )j,k κjk = 0, j = 0, 1, · · · , N − 1, (5.2)
dt
k=0
where
1
ωj = (d zj ∧ M d zj ), κjk = d zj ∧ K d zk ,
2
and zj = (pj , qj , vj , wj )T (j = 0, 1, · · · , N − 1).
2. Nonconservative multisymplectic Hamiltonian systems
Nonconservative multisymplectic Hamiltonian systems refer to those depending on
16.5 Some Generalizations 655
The corresponding four presymplectic forms associated to the time direction and three
space directions are respectively:
1 1
ω= d Z ∧ M (x, y, z) d Z, κx = d Z ∧ K d Z,
2 2
(5.5)
1 1
κy = d Z ∧ L d Z, κz = d Z ∧ N d Z.
2 2
656 16. Multisymplectic and Variational Integrators
Note that the time direction presymplectic form ω depends on the space variables
(x, y, z). We can also obtain the corresponding multisymplectic integrators[Che06a] .
3. Construction of multisymplectic integrators for modified equations
Consider the linear wave equation
Based on the two Hamiltonian formulations of (5.6) and using the hyperbolic func-
tions, various symplectic integrators were constructed in[QZ93] . By deriving the cor-
responding Lagrangians and their discrete counterparts, these symplectic integrators
were proved to be multisymplectic integrators for the modified versions of (5.6)
in[SQ00] .
Let us present an example. Using hyperbolic function tanh(x), we can obtain a
symplectic integrator for (5.6) of accuracy O(Δ t2s + Δ x2m ):
Δt Δt
uj+1
i − 2uji + uj−1
i = tanh 2s, tanh 2s, Δ (2m) (uj+1 i − 2uji + uij−1 ),
2 2
(5.7)
where
m−1 j
Δ x2 ∇+ ∇−
Δ (2m) = ∇+ ∇− (−1)j βj ,
4
j=0
where
[(j !)2 22j ]
βj =
[(2j + 1) ! (j + 1)]
and ∇+ and ∇− are forward and backward difference operators respectively.
For m = 2 and s = 2, the integrator (5.7) is a multisymplectic integrator of the
modified equation
Δ t2 Δ t4
utt = uxx − uxxxx − uxxxxxx . (5.8)
6 144
For other hyperbolic functions, we can obtain similar results.
and
1
z = (u, p, q)T , B = − eαt−βx (u2 + p2 − q 2 + αup + βuq),
2
T T
1 αt−βx 1 αt−βx 1 αt−βx 1
F = − e p, e u, 0 , G = e q, 0, − eαt−βx .
2 2 2 2
Similarly, we can develop multisymplectic dissipation integrators for the system (5.9)
which preserve a discrete version of the multisymplectic dissipation law (5.10).
5. Differential complex, methods and multisymplectic structure
Differential complexes have come to play an increasingly important role in numerical
analysis recently. In particular, discrete differential complexes are crucial in design-
ing stable finite element schemes[Arn02] . With regard to discrete differential forms, a
generic Hodge operator was introduced in[Hip02] . It was shown that most finite ele-
ment schemes emerge as its specializations. The connection between Veselov discrete
mechanics and finite element methods was first suggested in[MPS98] . Symplectic and
multisymplectic structures in simple finite element methods are explored in[GjLK04] . It
will be of particular significance to study the multisymplectic structure for the finite
element methods by using discrete differential complexes and in particular, discrete
Hodge operators[STMM07] . We will explore this issue in the future.
Bibliography
[AM04] U.M. Ascher and R.I. McLachlan: Multisymplectic box schemes and the Korteweg-de
Vries equation. Appl. Numer. Math., 39:55–269, (2004).
[Arn02] D.N. Arnold: Differential complexes and numerical stability. Plenary address deliv-
ered at ICM 2002. Beijing, China, (2002).
[BR01a] T. J. Bridges and S. Reich: Multi-symplectic integrators: numerical schemes for
Hamiltonian PDEs that conserve symplecticity. Physics Letters A, 284:184–193, (2001).
[BR01b] T.J. Bridges and S. Reich: Multi-symplectic spectral discretizations for the Zakharov-
Kuznetsov and shallow water equations. Physica D, 152:491–504, (2001).
[BR06] T. J. Bridges and S. Reich: Numerical methods for Hamiltonian PDEs. J. Phys. A:
Math. Gen., 39:5287–5320, (2006).
[Bri97] T. J. Bridges: Multi-symplectic structures and wave propagation. Math. Proc. Cam.
Phil. Soc., 121:147–190, (1997).
[Bri06] T. J. Bridges: Canonical multisymplectic structure on the total exterior algebra bundle.
Proc. R. Soc. Lond. A, 462:1531–1551, (2006).
[CGW03] J. B. Chen, H.Y. Guo, and K. Wu: Total variation in Hamiltonian formalism and
symplectic-energy integrators. J. of Math. Phys., 44:1688–1702, (2003).
[Che02] J. B. Chen: Total variation in discrete multisymplectic field theory and multisymplec-
tic energy momentum integrators. Letters in Mathematical Physics, 51:63–73, (2002).
[Che03] J. B. Chen: Multisymplectic geometry, local conservation laws and a multisymplectic
integrator for the Zakharov–Kuznetsov equation. Letters in Mathematical Physics, 63:115–
124, (2003).
[Che04a] J. B. Chen: Multisymplectic geometry for the seismic wave equation. Com-
mun.Theor. Phys., 41:561–566, (2004).
[Che04b] J. B. Chen: Multisymplectic Hamiltonian formulation for a one-way seismic wave
equation of high order approximation. Chin Phys. Lett., 21:37–39, (2004).
[Che05a] J. B. Chen: Multisymplectic geometry, local conservation laws and Fourier pseu-
dospectral discretization for the ”good” Boussinesq equation. Applied Mathematics and
Computation, 161:55–67, (2005).
[Che05b] J. B. Chen: A multisymplectic integrator for the periodic nonlinear Schrödinger
equation. Applied Mathematics and Computation, 170:1394–1417, (2005).
[Che05c] J. B. Chen: Variational formulation for multisymplectic Hamiltonian systems. Let-
ters in Mathematical Physics, 71:243–253, (2005).
[Che06a] J. B. Chen: A multisymplectic variational formulation for the nonlinear elastic wave
equation. Chin Phys. Lett., 23(2):320–323, (2006).
[Che06b] J. B. Chen: Symplectic and multisymplectic Fourier pseudospectral discretization
for the Klein-Gordon equation. Letters in Mathematical Physics, 75:293–305, (2006).
[Che07a] J. B. Chen: High order time discretization in seismic modeling. Geophysics,
72(5):SM115–SM122, (2007).
[Che07b] J. B. Chen: Modeling the scalar wave equation with Nyströn methods. Geophysics,
71(5):T158, (2007).
[Che07c] J. B. Chen: A multisymplectic pseudospectral method for seismic modeling. Applied
Mathematics and Computation, 186:1612–1616, (2007).
Bibliography 659
[CQ01a] J. B. Chen and M. Z. Qin: Multisymplectic fourier pseudospectral method for the
nonlinear Schrödinger equation. Electronic Transactions on Numerical Analysis, 12:193–
204, (2001).
[CQ02] J.-B. Chen and M. Z. Qin. A multisymplectic variational integrator for the nonlinear
Schrödinger equation. Numer. Meth. Part. Diff. Eq., 18:523–536, 2002.
[CQ03] J. B. Chen and M. Z. Qin: Multisymplectic composition integrators of high order. J.
Comput. Math., 21(5):647–656, (2003).
[CQT02] J. B. Chen, M. Z. Qin, and Y. F. Tang: Symplectic and multisymplectic methods for
the nonlinear Schrödinger equation. Computers Math. Applic., 43:1095–1106, (2002).
[CWQ09] J. Cai, Y. S. Wang, and Z. H. Qiao: Multisymplectic Preissman scheme for the
time-domain Maxwell’s equations. J. of Math. Phys., 50:033510, (2009).
[CYQ09] J. X. Cai, Y.S.Wang, and Z.H. Qiao: Multisymplectic Preissman scheme for the
time-domain Maxwell’s equations. J. of Math. Phys., 50:033510, (2009).
[CYWB06] J. X. Cai, Y.S.Wang, B. Wang, and B.Jiang: New multisymplectic self-adjoint
scheme and its composition for time-domain Maxwell’s equations. J. of Math. Phys.,
47:123508, (2006).
[Dah07] M. L. Dahlby: Geometrical integration of nonlinear wave equations. Master’s thesis,
Norwegian University, NTNU, Trondheim, (2007).
[Fen85] K. Feng: On difference schemes and symplectic geometry. In K. Feng, editor, Pro-
ceedings of the 1984 Beijing Symposium on Differential Geometry and Differential Equa-
tions, pages 42–58. Science Press, Beijing, (1985).
[FQ87] K. Feng and M. Z. Qin: The symplectic methods for the computation of Hamiltonian
equations. In Y. L. Zhu and B. Y. Guo, editors, Numerical Methods for Partial Differential
Equations, Lecture Notes in Mathematics 1297, pages 1–37. Springer, Berlin, (1987).
[GAR73] P.L. GARCIA: The Poincare–Cartan invariant in the calculus of variations symposia
mathematica. In in Convegno di Geometria Simplettica e Fisica Mathmatica XIV, pages
219–243. Academic Press, London, (1973).
[GjLK04] H.Y. Guo, X.M. Ji, Y.Q. Li, and K.Wu: symplectic, multisymplectic structure-
preserving in simple finite element method, Preprint arXiv: hep-th/0104151. (2004).
[Hip02] R. Hiptmair: Finite elements in computational electromagnetism. Acta Numerica,
11:237–339, (2002).
[HL04] J. Hong and C. Li: Multi-symplectic Runge–Kutta methods for nonlinear Dirac equa-
tions. J. of Comp. Phys., 211:448–472, (2004).
[HLHKA06] J. L. Hong, Y. Liu, H.Munthe-Kass, and Zanna A: On a multisymplectic scheme
for Schrödinger equations with variable coefficients. Appl. Numer. Math., 56:816–843,
(2006).
[HZQ03] L. Y Huang, W. P. Zeng, and M.Z. Qin: A new multi-symplectic scheme for nonlinear
“good” Boussinesq equation. J. Comput. Math., 21:703–714, (2003).
[JYJ06] B. Jiang, Y.S.Wang, and Cai J.X: New multisymplectic scheme for generalized
Kadomtsev-Petviashvili equation. J. of Math. Phys., 47:083503, (2006).
[KLX06] L. H. Kong, R. X. Liu, and Z.L. Xu: Numerical simulation interaction between
Schrödinger equation, and Klein–Gorden field by multi-symplecticic methods. Applied
Mathematics and Computation, 181:342–350, (2006).
[Lag88] J. L. Lagrange: Mécanique Analytique Blanchard, Paris, 5th edition, vol. 1, (1965).
[Lee82] T. D. Lee: Can time be a discrete dynamical variable? Phys.Lett.B, 122:217–220,
(1982).
[Lee87] T. D. Lee: Difference equations and conservation laws. J. Stat. Phys., 46:843–860,
(1987).
[LQ88] C.W. Li and M.Z. Qin: A symplectic difference scheme for the infinite dimensional
Hamiltonian system. J. Comput. Appl. Math, 6:164–174, (1988).
[LQ02] T. T. Liu and M. Z. Qin: Multisymplectic geometry and multisymplectic Preissman
scheme for the KP equation. J. of Math. Phys., 43:4060–4077, (2002).
660 Bibliography
[LQHD07] X. S. Liu, Y.Y. Qi, J. F. He, and P. Z. Ding: Recent progress in symplectic algo-
rithms for use in quantum systems. Communications in Computational Physics, 2(1):1–53,
(2007).
[MM05] K.W. Morton and D.F. Mayers: Numerical Solution of Partial Differential Equations:
an introduction. Cambridge University Press, Cambridge, Second edition, (2005).
[MPS98] J. E. Marsden, G.P. Patrick, and S. Shloller: Multi-symplectic geometry, variational
integrators, and nonlinear PDEs. Communications in Mathematical Physics, 199:351–395,
(1998).
[MPSM01] J. E. Marsden, S. Pekarsky, S. Shkoller, and M.West: Variational methods, multi-
symplectic geometry and continuum mechanics. J.Geom. Phys., 38:253–284, (2001).
[Olv86] P.J. Olver: Applications of Lie Groups to Differential Equations. Springer, New York,
(1986).
[Qin87] M. Z. Qin: A symplectic schemes for the Hamiltonian equations. J. Comput. Math.,
5:203–209, (1987).
[Qin90] M. Z. Qin: Multi-stage symplectic schemes of two kinds of Hamiltonian systems of
wave equations. Computers Math. Applic., 19:51–62, (1990).
[Qin97a] M. Z. Qin: A symplectic schemes for the PDEs. AMS/IP studies in Advanced Math-
emateics, 5:349–354, (1997).
[QZ92] M. Z. Qin and W. J. Zhu: Construction of higher order symplectic schemes by com-
position. Computing, 47:309–321, (1992).
[QZ93] M. Z. Qin and W. J. Zhu: Construction of symplectic scheme for wave equation via
hyperbolic functions sinh(x), cosh(x) and tanh(x). Computers Math. Applic., 26:1–11,
(1993).
[Rei00] S. Reich: Multi-symplectic Runge–Kutta collocation methods for Hamiltonian wave
equations. J. of Comp. Phys., 157:473–499, (2000).
[SHQ06] J. Q. Sun, W. Hua, and M. Z. Qin: New conservation scheme for the nonlinear
Schrodinger system. Applied Mathematics and Computation, 177:446–451, (2006).
[SMM04] J. Q. Sun, Z. Q. Ma, and M. Z. Qin: RKMK method of solving non-damping
LL equations for ferromagnet chain equations. Applied Mathematics and Computation,
157:407–424, (2004).
[SMQ06] J. Q. Sun, Z. Q. Ma, and M. Z. Qin: Simulation of envelope Rossby solution in pair
of cubic Schrodinger equations. Applied Mathematics and Computation, 183:946–952,
(2006).
[SNW92] J.C. Simo, N.Tarnow, and K.K. Wong: Exact energy-momentum conserving algo-
rithms and symplectic schemes for nonlinear dynamics. Comput. Methods Appl. Mech.
Engrg., 100:63–116, (1992).
[SQ00] Y. J. Sun and M.Z. Qin: Construction of multisymplectic schemes of any finite order
for modified wave equations. J. of Math. Phys., 41:7854–7868, (2000).
[SQ01] H. L. Su and M. Z. Qin: Multisymplectic Birkhoffian structure for PDEs with dissipa-
tion terms, arxiv:math.na 0302299, (2001).
[SQ03] H. Su and M. Z. Qin: Symplectic schemes for Birkhoffian system. Technical Report
arXiv: math-ph/0301001, (2003).
[SQ04] Y. J. Sun and M. Z. Qin: A multi-symplectic schemes for RLW eqution. J. Comput.
Math., 22:611–621, (2004).
[SQ05] H. Su and M. Z. Qin: Multisymplectic geometry method for Maxwell’s equations and
multisymplectic scheme. Technical Report arXiv. org math-ph/0302058, (2005).
[SQL06] J. Q. Sun, M.Z. Qin, and T.T. Liu: Total variation and multisymplectic structure for
the CNLS system. Commun.Theor. Phys., 46(2):966–975, (2006).
[SQS07] H. L. Su, M.Z. Qin, and R. Scherer: Multisymplectic geometry method for Maxwell’s
equations and multisymplectic scheme. Inter. J of Pure and Applied Math, 34(1):1–17,
(2007).
[SQWD09] J. Q. Sun, M. Z. Qin, H. Wei, and D. G. Dong: Numerical simulation of collision
behavior of optical solitons in birefingent fibres.
Commun Nonlinear Science and Numerical Simulation, 14:1259–1266, (2009).
Bibliography 661
Symbol Description
A, B Matrix A = {aij ∈ M (n)}
A∗ = A conjugate transpose of A
A , AT transpose of A
A J-orthogonal complement of A
A⊥ orthogonal complement of A
A = {Uλ , ϕλ } smooth atlas
Ad adjoint representation
Ad∗ coadjoint representation
adv adjoint vector field
ad∗v coadjoint vector field
Br (a) take a as the center of circle, r is the radius ball
Bk space consist of all exact k-form
Bk set of all k-boundaries
bk , bk Betti number
B(ρ), C(η), D(ζ) order conditions of Butcher.
C the complex numbers
Cn complex vector space of complex n-vector
Ck space of k-times differentiable functions
C∞ space of smooth functions
C(z) Casimir function
C k (M ) k-dimensional chain on M
i structure constant
Cjk
d exterior derivative, exterior differential operator
dxi basis differential 1-form
D total differential
det A determinant of matrix A
div divergence
deg ω (deg f )( deg P (x)) order of form (order of map) (order of polynomial)
Eτ Euler step-transient operator
e identity element of group
ex exponential function of x
ei , {ei , fj } basis, symplectic basis
eta phase of flow with vector field a
exp, Exp exponential map
F (t)f differential element of function f
664 Symbol
Symbol Description
F a field (usually R or C)
Fn vector space (over F) of n-vectors
f∗p differential of the map f in the p place
F(Rn ) a class of all differentiable function on Rn
g k (M ) set of all k-differential form on M
G group, Lie group
G2n,k M (2n, k) nonsingular equivalent class
g Lie algebra
g∗ dual to the Lie algebra
Gl(n), Gl(n, R), Gl(n, C) linear group on Rn ,(Cn )
gl(n) Lie algebra of n × n matrix
grad gradient
H K (M, R)(HK (M, R)) k-th cohomology (homology) group on M
H(p, q), H(z) Hamiltonian function
i including map
iX , contraction, interior product
I identity map
In , I2n identity matrix, standard Euclidean structure
id (Id) identity
im L image of map L
J momentum map
I2n , J4n symplectic structure
J"4n J"4n -symplectic structure
K K-symplectic structure
ker L kernel of mapping L
L[u] = L dx variation of L
LX Y, LX ω vector field Y , differential form ω of Lie derivative
M, N manifold
M (n, m, R) set of all real matrix with n-row and m-column
M (n, m, C) set of all complex matrix with n-row and m-column
M (n, R) set of all real matrix of order n × n on Rn
M (n, C) set of all complex matrix of order n × n on Cn
M (n, F) set of all matrix of order n × n on Fn
O(n), o(n) orthogonal group, orthogonal Lie algebra
O zero matrix
P p coordinate in momentum space
Q q coordinate in configuration space
p order of p
R real number
Rn n-dimensional real vector space
Rnp , Rnq momentum space, configuration space in Rn
RP n real projection space
r(t) order of t-tree
S symplectic transformation, S-transformation
h, s step of time
Symbol 665
Symbol Description
SL(n), SL(n, R), SL(n, C) special linear group, (real), (complex)
sl(n) Lie algebra of special linear group
SO(n) special orthogonal group
so(n) Lie algebra of special orthogonal group
Sp(2n) symplectic group, symplectic matrix
sp(2n) symplectic algebra, infinitesimal symplectic matrix
CSp(2n) conformal symplectic group
Sp(0) 0-class of symplectic matrix
Sp(I) I-class of symplectic matrix
Sp(II) II-class of symplectic matrix
Sp(III) III-class of symplectic matrix
Sp-diff or Sp-Diff symplectic diffeomorphism
TM tangent bundle
Tx M tangent space in place x
T ∗M cotangent bundle
Tx∗ M cotangent space in place of x
Sm symmetric group
u, v vector in Rn space
(U, ϕ), (V, ϕ) local coordinate
V vector space
Xp vector field in place p on manifold
ẋ, ẍ first, second, order derivative at x
x, xi x vector, coordinate component
y, y i y vector, coordinate component
X(M ) set of all tangent vector on M
XH Hamiltonian vector field
X (Rn ) set of all smooth vector field on Rn
A Bα
α
α= Darboux transformation
Cα Dα
Aα
Bα inverse of Darboux transformation
α−1 = α α
C D
δ variational derivative, codifferential operator
σ(t) symmetry of t-tree
γ(t) density of t-tree
δij Kronecker symbol
α(t) essential different labelings
Γf , Gf , gr (f ) graphic of f
Δt, τ, s step size of time
Δx step size of apace
θ differential 1-form
dθ exterior of differential 1-form
π T Rn −→ Rn projection T Rn to Rn
π −1 (x) = Tx Rn fiber in point x
666 Symbol
Symbol Description
ϕ∗ ω (ϕ∗ ω) pull back of differential form (push-forward)
ϕ∗ f (ϕ∗ f ) pull back of function(push-forward)
ϕ∗ Y (ϕ∗ Y ) pull back of vector field (push-forward)
× product
∧ exterior product
Λk (Rn ) k-th exterior bundle over Rn
Λn Lagrangian subspace
Λn (K) K-Lagrangian subspace
f Z f transverse to Z
f p Z f in the p transverse to Z
Ω standard symplectic structure
Ω# lift of mapping Ω# (z1 )(z2 ) = Ω−1 (z1 , z2 )
Ωb down mapping Ωb (z1 )(z2 ) = Ω(z1 , z2 )
Ωk (Rn ), Ω0 (Rn ) = C ∞ (Rn ) k-differential form on Rn
∂
∂xI
or ∂xi partial derivative with respect to xi
∂ boundary operator
(×f rotation
(
·f divergence
pds boundary integral
ω integral of differential form
∅ empty set
⊗ tensor product
∩ set-theoretic intersection
∪ set-theoretic union
⊂ inclusion
∈ element of
element of
◦ f ◦ g = f (g) composition
/ division
∈ not element of
∀ for
homomorphism
approximate
∼
= similarly
≡ identity
:= definition
∼ corresponding, equivalent, congruent relation
c
∼ conjugate congruent
−→ mapping
=⇒ extrusion
⇐⇒ extrusion mutually
n
n! binomial coefficient
Cnk = =
k k!(n − k)!
Symbol 667
Symbol Description
* +
n n!
= multinomial coefficients.
k1 , k 2 , · · · , k r k1 !k2 ! · · · , kr !
where k1 + k2 + · · · + kr = n
(a, b) open interval
[a, b] closed interval
[u, w] Lie bracket
[A, B] matrix commutator
[F, H] Poisson bracket
(u, v) inner product, Euclidean inner product
[U, V ] symplectic inner product
B norm of matrix
◦ direct sum
U +V
symplectic direct sum
P1 "P
2 inner product
,
Poisson bracket
{ϕ, φ}
vector a orthogonal to b (Euclidean)
a⊥b
ab vector a symplectic orthogonal to b
1N (x) = x identity function
Index
A C
A(α)-stability, 550 calculate the formal energy, 267
a*–linear differential operator, 407 canonical equation, 170
a∗ –Jacobian matrix, 407 canonical forms under orthogonal
A-stability, 550 transformation, 134
ABC flow, 446 canonical reductions of bilinear forms, 128
action functional of Lagrangian density, 643 canonical transformation, 172, 188
Ad*-equivariant, 503 Cartan form, 644
adjoint integrator, 374 Cartan’s Magic formula, 106
adjoint method, 372 Casimir function, 501
all polynomials is symplectically separability Cayley transformation, 193
in R2n , 207 centered Euler method, 416
alternative canonical forms, 130 centered Euler scheme, 192, 200, 231
angular momentum in body description, 505 chains, 91
angular momentum in space description, 505 characteristic equations, 477
angular momentum-preserving schemes for chart, 40
rigid body, 525 Chebyshev spectral method, 508
angular velocity in body description, 505 classical Stokes theorem, 98
angular velocity in space description, 505 closed form, 84
anti-symmetric product, 117 closed nondegenerate differential 2-form, 165
atlas, 40 coadjoint orbits, 505
automorphism, 39 coclosed form, 90
autonomous Birkhoff’s equations, 618 codifferential operator, 89
coefficient B-series for centered Euler
scheme, 418
B coefficient B-series for exact solution, 418
coefficient B-series for explicit Euler scheme,
B-series, 417 418
B-stability, 550 coefficient B-series for implicit Euler scheme,
backward error analysis, 432 418
base of tangent space, 45 coefficient B-series for R–K method, 418
BCH formula, 380, 413 coefficient B-series for trapezoidal scheme,
Betti numbers, 99 418
bijective, 39 coefficients can be determined recursively,
bilinear antisymmetric form, 188 233
binary forms, 116 coexact form, 90
Birkhoffian system, 618 cohomology space, 98
black (fat )vertex, 309 coisotropic subspace, 138
boundary of chains, 92 commutativity of generator maps, 261
Butcher tableau, 278 commutator, 124, 179
670 Index
commutator of two vector fields, 100 constructing s-scheme by Poincaré type g.f.,
comparison order conditions between 229
symplectic R–K (R–K–N) method, 302 constructing s-scheme via 1st kind g.f., 227
comparison order conditions P–R–K method construction of volume-preserving schemes
and symplectic P–R–K method, 318, 319, via g.f., 464
333 contact 1-form, 480
compatible of two local coordinate systems, contact algorithm, 483
40 contact algorithm–C, 493
complete non-integrability, 477 contact algorithm–P , 492
complexifiable, 124 contact algorithm–Q, 492
complexification of real vector space and real contact difference schemes, 492
linear transformation, 123 contact dynamical systems, 477
composition laws, 419 contact element, 482
composition of centered Euler scheme, 372 contact generating function, 487
composition of trapezoid scheme, 365 contact geometry, 477
composition scheme is not A-stable, 389 contact Hamiltonian, 483, 492
compositional property of Lie series, 379 contact map, 486
condition for centered Euler to be volume- contact structure, 477, 481
preserving, 444 contact transformation, 483
condition of symplectic P–R–K method, 303 contactization of conic symplectic maps, 487
condition of variational self-adjointness, 619 contraction, 105
configuration space, 188 convergence of symplectic difference
conformally K-symplectic group schemes, 239
CSp(K, n, F), 120 coordinate Lagrangian subspaces, 147
conformally canonical transformation, 173, coordinate of tangent vector, 45
182 coordinate subspaces, 139
conformally Hermitian, 117 cotangent bundle, 76, 249
conformally identical, 114 cotangent vector, 76
conformally orthogonal group CO(S, n, F), cycle, 93
120
conformally symmetric, 114 D
conformally symplectic group CSp(2n), 144
conformally unitary group CU (H, n, C), Darboux matrix, 231, 600
120 Darboux theorem, 168, 190
congruence canonical forms of conformally Darboux transformation, 249
symmetric, 130 De Rham theorem, 99
congruence canonical forms of Hermitian decomposed theorem of symplectic matrix,
matrices, 130 155
congruent reductions, 129 decompositions of source-free vector fields,
conic function, 484 452
conic Hamiltonian vector fields, 488 definition of symplectic for LMM, 356
conic map, 484 density of tree γ(t), 294
conic symplectic, 484 diagonal formal flow, 415
conic symplectic map, 484 diagonal Padé approximant, 194
conic transformation, 488 diagonally implicit method, 284
conservation Laws, 234 diagonally implicit symplectic R–K method,
conservation of spatial angular momentum 284
theorem, 506 diffeomorphism, 39, 102, 126, 188
constrained Hamiltonian algorithm, 537 diffeomorphism group, 102
construction of the difference schemes via differentiable manifold, 40
generating function, 213 differentiable manifold structure, 40
construct volume-preserving difference differentiable mapping, 41
schemes, 454 differentiable mapping, differential concept,
constructing s-scheme by 2nd kind g.f., 227 43
Index 671
obstruction, 450 Q
one-form (1-form), 66
one-leg weighted Euler schemes, 231 quadratic bilinear form, 115
one-parameter group of canonical maps, 221 quaternion form, 524
operation ∧, 65
optimization Method, 603 R
orbit-preserving schemes, 527
order conditions for symplectic R–K–N Radau I A, 279
method, 319 Radau IA-IA, 471
orientable differentiable manifold, 59 Radau II A, 280
orientable vector spaces, 59 Radau IIA-IIA, 472
Radau polynomial, 279
orthogonal group O(n, F), 119
rational fraction, 200
real representation of complex vector space,
P 121
reduction method, 540
P–R–K method, 302
reflective polynomial, 158
Padé approximation, 193
regular submanifold, 51, 53
Padé approximation table, 196
relationship between rooted tree and
partitions and skeletons, 418 elementary differential, 293
Pfaffian theorem, 118 resonant, 568
phase flow, 102, 221, 408 revertible approximations, 450
phase flow of contact system, 483 Riemann structure, 167
phase flow- etF , 235 right translation, 503
phase space, 102 rigid body in Euclidean space, 523
phase-area conservation law, 189 Rodrigue formula, 543
Poincaré lemma, 85, 220, 222 root isomorphism, 298
Poincaré transformation, 250 rooted n-tree, 299
Poincaré’s generating function and H.J. rooted P -tree, 309
equation, 223 rooted S-tree, 321
Poisson bracket, 177, 192, 499 rooted 3-tree, 298
Poisson manifold, 499 rooted labeled n-tree ρλτ , 297
Poisson mapping, 500 rooted labeled P -tree, 309
Poisson scheme, 508 rooted labeled S-tree, 321
Poisson system, 500 rooted labeled 3-tree, 298
Poisson theorem, 179 rooted labeled trees, 298
postprocessed vector field, 432
Preissman integrator, 646 S
preprocessed vector field integrators, 432
preserve all quadratic first integrals of system, S-graph, 321
236 S-orthogonal group, 119
preserve angular momentum pT Bq, 236 S-tree, 321
preserving the contact structure, 483 scalar product, 117
presymplectic form, 645 section of tangent bundle, 62
presymplectic forms, 605 self-adjoint integrator, 376
product of cotangent bundles, 249 self-adjoint method, 372
Index 675
semi-autonomous Birkhoff’s equation, 618 symplectic conditions for R–K method, 281
separable Hamiltonian system, 202 symplectic explicit R–K–N method
separable systems for source-free systems, (non-redundant 5-stage fifth order), 331
447 symplectic form, 118
sesquilinear form, 116 symplectic geometry, 165, 188
simplify symplectic R–K conditions, 300 symplectic group, 188
simplifying condition of R–K method, 279 symplectic group Sp(2n), 144
Sm(2n)matrices, 600 symplectic group Sp(2n, F ), 119
small twist mappings, 558 symplectic invariant algorithms, 235
some theorems about Sp(2n), 151 symplectic leave, 505
sons of the root, 297 symplectic LMM for linear Hamiltonian
source-free system, 443, 449, 467 systems, 348
Sp(2n) matrices, 600 symplectic manifold, 165
SpD2n the totality of symplectic operators, symplectic map, 220
232 symplectic mapping, 215
SpD2n the set of symplectic transformations, symplectic matrix, 189
601 symplectic operators near identity, 232
special linear group SL(n, F ), 119 symplectic pair, 217
special separable source-free systems, 458 symplectic R–K method, 277, 279
special type Sp2n (I), 150 symplectic R–K–N method, 319
special type Sp2n (II), 151 symplectic R–K–N method (3-stage and 4-th
special type Sp2n (III), 151 order), 323
special type Sp2n (I, II), 151 symplectic schemes for Birkhoffian Systems,
special types of Sp(2n), 148 625
stability analysis for composition scheme, symplectic schemes for nonautonomous
388 system, 244
standard antisymmetric matrix, 192 symplectic space, 137
standard symplectic structure, 169, 188, 249 symplectic structure, 137, 165, 215, 477
star operators, 88 symplectic structure for trapezoidal scheme,
step size resonance, 568 202
step transition, 415 symplectic structure in product space, 215
step-forward operator, 240 symplectic subspace, 137
Stokes theorem, 93 symplectic-energy integrator, 596, 602
structure-stability, 551 symplectic-energy-momentum, 581
subalgebra of a Lie algebra, 179 symplectically separable Hamiltonian
submanifold, 46 systems, 205
submersion, 51 symplectization of contact space, 487
substitution law, 432 symplified order conditions for symplectic
superfluous trees, 298 R–K–N method, 327
surjective, 39 symplified order conditions of explicit
Sylvester’s law of inertia, 132 symplectic R–K method, 307
Symm(2n) the set of symmetric
transformations, 601 T
symm(2n) the totality of symmetric
operators, 232 table of coefficient ω(τ ) for trees of order
symmetric operators near nullity, 232 5, 435
symmetric pair, 216 table of coefficients σ(τ ), γ(τ ), b̆(τ ), and b(τ ),
symmetric product, 117 434
symmetrical composition, 376 table of composition laws for the trees of
symmetry of tree σ(t), 294 order ≤ 4, 436
symplectic algebra, 216 table of substitution law ∗ defined in for the
symplectic algorithms as small twist trees of order ≤ 5, 437
mappings, 560 tangent bundle, 56
symplectic basis, 145 tangent mapping, 58
676 Index
tangent space, 44
tangent vector, 43
the elementary differential, 291
the inverse function to exp, 126
the order of tree r(t), 294
time-dependent gradient map, 221
topological manifold, 40
total variation for Lagrangian mechanics, 583
total variation in Hamiltonian mechanics, 593
transversal, 54, 140, 143
transversal Lagrangian subspaces, 148
transversality condition, 181, 213, 221, 225,
227, 250, 251, 460, 623
trapezoidal method, 416
trapezoidal scheme, 201
tree, 298
trivial tangent bundle, 57
truncation, 233
two-forms (2-forms), 66
U
Unitary group U (n, C), 119
Unitary product, 118
V
variational integrators, 651
variational principle in Hamiltonian
mechanics, 591
vector field, 62
vertical vector field, 582
Veselov–Moser algorithm, 539
volume-preserving 2-Stage P–R–K methods,
471
volume-preserving P-R–K method, 467
volume-preserving R–K method, 467
volume-preserving schemes, 444
W
W -transformation, 304, 470
white (meagre) vertex, 309
Witt theorem, 132
X
X-matrix, 305