3.1 Linear, constant coefficient problems
We consider an evolution equation on
-dimensional space of the following form:
Here,
is the state vector, and
its partial derivative with respect to
. Next, the
’s denote complex,
matrices where
denotes a multi-index with
components
and
. Finally,
denotes the partial derivative
operator
of order
, where
. Here are a few representative examples:
Example 1. The advection equation
with speed
in the negative
direction.
Example 2. The heat equation
, where
denotes the Laplace operator.
Example 3. The Schrödinger equation
.
Example 4. The wave equation
, which can be cast into the form of Eq. (3.1),
We can find solutions of Eq. (3.1) by Fourier transformation in space,
Applied to Eq. (3.1) this yields the system of linear ordinary differential equations
for each wave vector
where
, called the symbol of the differential operator
, is
defined as
The solution of Eq. (3.4) is given by
where
is determined by the initial data for
at
. Therefore, the formal solution of the
Cauchy problem
with given initial data
for
at
is
where
.
3.1.1 Well-posedness
At this point we have to ask ourselves if expression (3.9) makes sense. In fact, we do not expect the integral
to converge in general. Even if
is smooth and decays rapidly to zero as
we could still have
problems if
diverges as
. One simple, but very restrictive, possibility to
control this problem is to limit ourselves to initial data
in the class
of functions, which
are the Fourier transform of a
-function with compact support, i.e.,
, where
A function in this space is real analytic and decays faster than any polynomial as
.
If
the integral in Eq. (3.9) is well-defined and we obtain a solution of the Cauchy problem (3.7,
3.8), which, for each
lies in this space. However, this possibility suffers from several unwanted
features:
- The space of admissible initial data is very restrictive. Indeed, since
is necessarily
analytic it is not possible to consider nontrivial data with, say, compact support, and study
the propagation of the support for such data.
- For fixed
, the solution may grow without bound when perturbations with arbitrarily
small amplitude but higher and higher frequency components are considered. Such an effect is
illustrated in Example 6 below.
- The function space
does not seem to be useful as a solution space when considering linear
variable coefficient or quasilinear problems, since, for such problems, the different
modes do
not decouple from each other. Hence, mode coupling can lead to components with arbitrarily
high frequencies.
For these reasons, it is desirable to consider initial data of a more general class than
. For this, we need to
control the growth of
. This is captured in the following
Definition 1. The Cauchy problem (3.7, 3.8) is called well posed if there are constants
and
such that
The importance of this definition relies on the property that for each fixed time
the norm
of the propagator is bounded by the constant
, which is independent of the wave
vector
. The definition does not state anything about the growth of the solution with time other that
this growth is bounded by an exponential. In this sense, unless one can choose
or
arbitrarily small, well-posedness is not a statement about the stability in time, but rather about stability
with respect to mode fluctuations.
Let us illustrate the meaning of Definition 1 with a few examples:
Example 5. The heat equation
.
Fourier transformation converts this equation into
. Hence, the symbol is
and
. The problem is well posed.
Example 6. The backwards heat equation
.
In this case the symbol is
, and
. In contrast to the previous case,
exhibits exponential frequency-dependent growth for each fixed
and the problem is not well posed.
Notice that small initial perturbations with large
are amplified by a factor that becomes larger and
larger as
increases. Therefore, after an arbitrarily small time, the solution is contaminated by
high-frequency modes.
Example 7. The Schrödinger equation
.
In this case we have
and
. The problem is well posed. Furthermore, the
evolution is unitary, and we can evolve forward and backwards in time. When compared to the previous
example, it is the factor
in front of the Laplace operator that saves the situation and allows the
evolution backwards in time.
Example 8. The one-dimensional wave equation written in first-order form,
The symbol is
. Since the matrix
is symmetric and has eigenvalues
, there exists an
orthogonal transformation
such that
Therefore,
, and the problem is well posed.
Example 9. Perturb the previous problem by a lower-order term,
The symbol is
, and
. The problem is well posed, even though the
solution grows exponentially in time if
.
More generally one can show (see Theorem 2.1.2 in [259
]):
By considering the eigenvalues of the symbol
we obtain the following simple necessary condition
for well-posedness:
Proof. Suppose
is an eigenvalue of
with corresponding eigenvector
,
. Then,
if the problem is well posed,
for all

, which implies that

for all

, and hence

. □
Although the Petrovskii condition is a very simple necessary condition, we stress that it is not sufficient
in general. Counterexamples are first-order systems, which are weakly, but not strongly, hyperbolic; see
Example 10 below.
3.1.2 Extension of solutions
Now that we have defined and illustrated the notion of well-posedness, let us see how it can be used to solve
the Cauchy problem (3.7, 3.8) for initial data more general than in
. Suppose first that
, as
before. Then, if the problem is well posed, Parseval’s identities imply that the solution (3.9) must satisfy
Therefore, the
-solution satisfies the following estimate
for all
. This estimate is important because it allows us to extend the solution to the much
larger space
. This extension is defined in the following way: let
. Since
is dense in
there exists a sequence
in
such that
.
Therefore, if the problem is well posed, it follows from the estimate (3.18) that the corresponding
solutions
defined by Eq. (3.9) form a Cauchy-sequence in
, and we can define
where the limit exists in the
sense. The linear map
satisfies the
following properties:
is the identity map.
for all
.
- For
,
is the unique solution to the Cauchy problem (3.7, 3.8).
for all
and all
.
The family
is called a semi-group on
. In general,
cannot be extended to
negative
as the example of the backwards heat equation, Example 6, shows.
For
the function
is called a weak solution of the Cauchy
problem (3.7, 3.8). It can also be constructed in an abstract way by using the Fourier–Plancharel
operator
. If the problem is well posed, then for each
and
the map
defines an
-function, and, hence, we can define
According to Duhamel’s principle, the semi-group
can also be used to construct weak
solutions of the inhomogeneous problem,
where
,
is continuous:
For a discussion on semi-groups in a more general context see Section 3.4.
3.1.3 Algebraic characterization
In order to extend the solution concept to initial data more general than analytic, we have introduced the
concept of well-posedness in Definition 1. However, given a symbol
, it is not always a simple task
to determine whether or not constants
and
exist such that
for all
and
. Fortunately, the matrix theorem by Kreiss [257] provides necessary and sufficient
conditions on the symbol
for well-posedness.
A generalization and complete proof of this theorem can be found in [259
]. However, let us show
here the implication (ii)
(i) since it illustrates the concept of energy estimates, which
will be used quite often throughout this review (see Section 3.2.3 below for a more general
discussion of these estimates). Hence, let
be a family of
Hermitian matrices
satisfying the condition (3.25). Let
and
be fixed, and define
for
. Then we have the following estimate for the “energy” density
,
which implies the differential inequality
Integrating, we find
which implies the inequality (3.24) with
.
3.1.4 First-order systems
Many systems in physics, like Maxwell’s equations, the Dirac equation, and certain formulation of Einstein’s
equations are described by first-order partial-differential equations (PDEs). In fact, even systems, which are
given by a higher-order PDE, can be reduced to first order at the cost of introducing new variables, and
possibly also new constraints. Therefore, let us specialize the above results to a first-order linear problem of
the form
where
are complex
matrices. We split
into its
principal symbol,
, and the lower-order term
. The principal part is
the one that dominates for large
and hence the one that turns out to be important for
well-posedness. Notice that
depends linearly on
. With these observations in mind we
note:
- A necessary condition for the problem to be well posed is that for each
with
the
symbol
is diagonalizable and has only purely imaginary eigenvalues. To see this, we require
the inequality
for all
and
,
, replace
by
, and take the limit
, which
yields
for all
with
. Therefore, there must exist for each such
a complex
matrix
such that
, where
is a
diagonal real matrix (cf. Lemma 1).
- In this case the family of Hermitian
matrices
satisfies
for all
with
.
- However, in order to obtain the energy estimate, one also needs the condition
,
that is,
must be uniformly bounded and positive. This follows automatically if
depends continuously on
, since
varies over the
-dimensional unit sphere, which is
compact.
In turn, it follows that
depends continuously on
if
does. However, although this
may hold in many situations, continuous dependence of
on
cannot always be established;
see Example 12 for a counterexample.
These observations motivate the following three notions of hyperbolicity, each of them being a stronger
condition than the previous one:
The matrix theorem implies the following statements:
- Strongly and symmetric hyperbolic systems give rise to a well-posed Cauchy problem. According to
Theorem 1, their principal symbol satisfies
and this property is stable with respect to lower-order perturbations,
The last inequality can be proven by applying Duhamel’s formula (3.23) to the function
, which satisfies
with
. The solution
formula (3.23) then gives
, which yields
upon
integration.
- As we have anticipated above, a necessary condition for well-posedness is the existence of a
complex
matrix
for each
on the unit sphere, which brings the
principal symbol
into diagonal, purely imaginary form. If, furthermore,
can
be chosen such that
and
are uniformly bounded for all
,
then
satisfies the conditions (3.31) for strong hyperbolicity.
If the system is well posed, Theorem 2.4.1 in [259
] shows that it is always possible to
construct a symmetrizer
satisfying the conditions (3.31) in this manner, and hence,
strong hyperbolicity is also a necessary condition for well-posedness. The symmetrizer
construction
is useful for applications, since
is easily
constructed from the eigenvectors and
from the eigenfields of the principal symbol; see
Example 15.
- Weakly hyperbolic systems are not well posed in general because
might exhibit polynomial
growth in
. Although one might consider such polynomial growth as acceptable, such
systems are unstable with respect to lower-order perturbations. As the next example
shows, it is possible that
grows exponentially in
if the system is weakly
hyperbolic.
Example 10. Consider the weakly hyperbolic system [259
]
with
a parameter. The principal symbol is
and
Using the tools described in Section 2 we find for the norm
which is approximately equal to
for large
. Hence, the solutions to Eq. (3.35) contain modes,
which grow linearly in
for large
when
, i.e., when there are no lower-order
terms.
However, when
, the eigenvalues of
are
which, for large
has real part
. The eigenvalue with positive real part gives rise
to solutions, which, for fixed
, grow exponentially in
.
Example 11. For the system [353
],
the principal symbol,
, is diagonalizable for all vectors
except for those with
. In particular,
is diagonalizable for
and
. This shows that in general, it is not sufficient to check that the
matrices
,
,…,
alone are diagonalizable and have real eigenvalues; one has to consider all possible linear combinations
with
.
Example 12. Next, we present a system for which the eigenvectors of the principal symbol cannot be
chosen to be continuous functions of
:
The principal symbol
has eigenvalues
and for
the corresponding eigenprojectors are
When
the two eigenvalues fall together, and
converges to the zero matrix.
However, it is not possible to continuously extend
to
. For example,
for positive
. Therefore, any choice for the matrix
, which diagonalizes
,
must be discontinuous at
since the columns of
are the eigenvectors of
.
Of course,
is symmetric and so
can be chosen to be unitary, which yields the trivial
symmetrizer
. Therefore, the system is symmetric hyperbolic and yields a well-posed Cauchy
problem; however, this example shows that it is not always possible to choose
as a continuous
function of
.
Example 13. Consider the Klein–Gordon equation
in two spatial dimensions, where
is a parameter, which is proportional to the mass of
the field
. Introducing the variables
we obtain the first-order system
The matrix coefficients in front of
and
are symmetric; hence the system is symmetric hyperbolic with trivial symmetrizer
.
The corresponding Cauchy problem is well posed. However, a problem with this first-order system is that it
is only equivalent to the original, second-order equation (3.43) if the constraints
and
are satisfied.
An alternative symmetric hyperbolic first-order reduction of the Klein–Gordon equation, which does
not require the introduction of constraints, is the Dirac equation in two spatial dimensions,
This system implies the Klein–Gordon equation (3.43) for either of the two components of
.
Yet another way of reducing second-order equations to first-order ones without introducing constraints
will be discussed in Section 3.1.5.
Example 14. In terms of the electric and magnetic fields
, Maxwell’s evolution equations,
constitute a symmetric hyperbolic system. Here,
is the current density and
and
denote the
nabla operator and the vector product, respectively. The principal symbol is
and a symmetrizer is given by the physical energy density,
in other words,
is trivial. The constraints
and
propagate as a consequence
of Eqs. (3.46, 3.47), provided that the continuity equation holds:
,
.
Example 15. There are many alternative ways to write Maxwell’s equations. The following
system [353
, 287
] was originally motivated by an analogy with certain parametrized first-order hyperbolic
formulations of the Einstein equations, and provides an example of a system that can be symmetric,
strongly, weakly or not hyperbolic at all, depending on the parameter values. Using the Einstein summation
convention, the evolution system in vacuum has the form
where
and
,
, represent the Cartesian components of the electric field and
the gradient of the magnetic potential
, respectively, and where the real parameters
and
determine the dynamics of the constraint hypersurface defined by
and
.
In order to analyze under which conditions on
and
the system (3.50, 3.51) is strongly
hyperbolic we consider the corresponding symbol,
Decomposing
and
into components parallel and orthogonal to
,
where in terms of the projector
orthogonal to
we have defined
,
and
,
,
,
, and
,
we can write the eigenvalue problem
as
It follows that
is diagonalizable with purely complex eigenvalues if and only if
. However,
in order to show that in this case the system is strongly hyperbolic one still needs to construct a bounded
symmetrizer
. For this, we set
and diagonalize
with
and
Then, the quadratic form associated with the symmetrizer is
and
is smooth in
. Therefore, the system is indeed strongly hyperbolic for
.
In order to analyze under which conditions the system is symmetric hyperbolic we notice that because of
rotational and parity invariance the most general
-independent symmetrizer must have the form
with strictly positive constants
,
,
and
, where
denotes the
symmetric, trace-free part of
and
its trace. Then,
For
to be a symmetrizer, the expression on the right-hand side must be purely imaginary. This
is the case if and only if
,
and
. Since
,
,
and
are positive, these equalities can be satisfied if and only if
and
. Therefore, if either
and
are both positive or
and
are both negative and
or
, then the system (3.50, 3.51) is strongly but not symmetric hyperbolic.
3.1.5 Second-order systems
An important class of systems in physics are wave problems. In the linear, constant coefficient case, they are
described by an equation of the form
where
is the state vector, and
denote complex
matrices. In order to apply the theory described so far, we reduce this equation
to a system that is first order in time. This is achieved by introducing the new variable
.
With this redefinition one obtains a system of the form (3.1) with
and
Now we could apply the matrix theorem, Theorem 1, to the corresponding symbol
and analyze
under which conditions on the matrix coefficients
the Cauchy problem is well posed.
However, since our problem originates from a second-order equation, it is convenient to rewrite the symbol
in a slightly different way: instead of taking the Fourier transform of
and
directly, we multiply
by
and write the symbol in terms of the variable
. Then, the
-norm of
controls, through Parseval’s identity, the
-norms of the first partial derivatives of
, as is
the case for the usual energies for second-order systems. In terms of
the system reads
in Fourier space, where
with
. As for first-order systems, we can split
into its principal part,
which dominates for
, and the remaining, lower-order terms. Because of the homogeneity of
in
we can restrict ourselves to values of
on the unit sphere, like for
first-order systems. Then, it follows as a consequence of the matrix theorem that the problem is
well posed if and only if there exists a symmetrizer
and a constant
satisfying
for all such
. Necessary and sufficient conditions under which such a symmetrizer exists have been given
in [261
] for the particular case in which the mixed–second-order derivative term in Eq. (3.56) vanishes; that
is, when
. This result can be generalized in a straightforward manner to the case where the
matrices
are proportional to the identity:
Proof. Since for
the advection term
commutes with any Hermitian matrix
, it is sufficient to prove the theorem for
, in which case the principal symbol reduces to
We write the symmetrizer

in the following block form,
where

,

and

are complex

matrices, the first two being Hermitian.
Then,
Now, suppose

satisfies the conditions (
3.63). Then, choosing

,

and

we find that

. Furthermore,

and

where
is finite because

is continuous in

and

. Therefore,

is a symmetrizer for

,
and the problem is well posed.
Conversely, suppose that the problem is well posed with symmetrizer
. Then, the vanishing of
yields the conditions
and the
conditions (3.63) are satisfied for
. □
Remark: The conditions (3.63) imply that
is symmetric and positive with respect to the scalar
product defined by
. Hence it is diagonalizable, and all its eigenvalues are positive. A practical way of
finding
is to construct
, which diagonalizes
,
with
diagonal and positive. Then,
is the candidate for satisfying the
conditions (3.63).
Let us give some examples and applications:
Example 16. The Klein–Gordon equation
on flat spacetime. In this case,
and
, and
trivially satisfies the conditions of Theorem 2.
Example 17. In anticipation of the following Section 3.2, where linear problems with variable coefficients
are treated, let us generalize the previous example on a curved spacetime
. We assume that
is globally hyperbolic such that it can be foliated by space-like hypersurfaces
. In the ADM
decomposition, the metric in adapted coordinates assumes the form
with
the lapse,
the shift vector, which is tangent to
, and
the
induced three-metric on the spacelike hypersurfaces
. The inverse of the metric is given by
where
are the components of the inverse three-metric. The Klein–Gordon equation on
is
which, in the constant coefficient case, has the form of Eq. (3.56) with
Hence,
, and the conditions of Theorem 2 are satisfied with
since
and
is symmetric positive definite.