In contrast to previous notes, in this set of notes we shall focus exclusively on Fourier analysis in the one-dimensional setting for simplicity of notation, although all of the results here have natural extensions to higher dimensions. Depending on the physical context, one can view the physical domain
as representing either space or time; we will mostly think in terms of the former interpretation, even though the standard terminology of “time-frequency analysis”, which we will make more prominent use of in later notes, clearly originates from the latter.
In previous notes we have often performed various localisations in either physical space or Fourier space , for instance in order to take advantage of the uncertainty principle. One can formalise these operations in terms of the functional calculus of two basic operations on Schwartz functions
, the position operator
defined by
and the momentum operator , defined by
(The terminology comes from quantum mechanics, where it is customary to also insert a small constant on the right-hand side of (1) in accordance with de Broglie’s law. Such a normalisation is also used in several branches of mathematics, most notably semiclassical analysis and microlocal analysis, where it becomes profitable to consider the semiclassical limit
, but we will not emphasise this perspective here.) The momentum operator can be viewed as the counterpart to the position operator, but in frequency space instead of physical space, since we have the standard identity
for any and
. We observe that both operators
are formally self-adjoint in the sense that
for all , where we use the
Hermitian inner product
Clearly, for any polynomial of one real variable
(with complex coefficients), the operator
is given by the spatial multiplier operator
and similarly the operator is given by the Fourier multiplier operator
Inspired by this, if is any smooth function that obeys the derivative bounds
for all and
(that is to say, all derivatives of
grow at most polynomially), then we can define the spatial multiplier operator
by the formula
one can easily verify from several applications of the Leibniz rule that maps Schwartz functions to Schwartz functions. We refer to
as the symbol of this spatial multiplier operator. In a similar fashion, we define the Fourier multiplier operator
associated to the symbol
by the formula
For instance, any constant coefficient linear differential operators can be written in this notation as
however there are many Fourier multiplier operators that are not of this form, such as fractional derivative operators for non-integer values of
, which is a Fourier multiplier operator with symbol
. It is also very common to use spatial cutoffs
and Fourier cutoffs
for various bump functions
to localise functions in either space or frequency; we have seen several examples of such cutoffs in action in previous notes (often in the higher dimensional setting
).
We observe that the maps and
are ring homomorphisms, thus for instance
and
for any obeying the derivative bounds (2); also
is formally adjoint to
in the sense that
for , and similarly for
and
. One can interpret these facts as part of the functional calculus of the operators
, which can be interpreted as densely defined self-adjoint operators on
. However, in this set of notes we will not develop the spectral theory necessary in order to fully set out this functional calculus rigorously.
In the field of PDE and ODE, it is also very common to study variable coefficient linear differential operators
where the are now functions of the spatial variable
obeying the derivative bounds (2). A simple example is the quantum harmonic oscillator Hamiltonian
. One can rewrite this operator in our notation as
and so it is natural to interpret this operator as a combination of both the position operator
and the momentum operator
, where the symbol
this operator is the function
Indeed, from the Fourier inversion formula
for any we have
and hence on multiplying by and summing we have
Inspired by this, we can introduce the Kohn-Nirenberg quantisation by defining the operator by the formula
whenever and
is any smooth function obeying the derivative bounds
for all and
(note carefully that the exponent in
on the right-hand side is required to be uniform in
). This quantisation clearly generalises both the spatial multiplier operators
and the Fourier multiplier operators
defined earlier, which correspond to the cases when the symbol
is a function of
only or
only respectively. Thus we have combined the physical space
and the frequency space
into a single domain, known as phase space
. The term “time-frequency analysis” encompasses analysis based on decompositions and other manipulations of phase space, in much the same way that “Fourier analysis” encompasses analysis based on decompositions and other manipulations of frequency space. We remark that the Kohn-Nirenberg quantization is not the only choice of quantization one could use; see Remark 19 below.
In principle, the quantisations are potentially very useful for such tasks as inverting variable coefficient linear operators, or to localize a function simultaneously in physical and Fourier space. However, a fundamental difficulty arises: map from symbols
to operators
is now no longer a ring homomorphism, in particular
in general. Fundamentally, this is due to the fact that pointwise multiplication of symbols is a commutative operation, whereas the composition of operators such as and
does not necessarily commute. This lack of commutativity can be measured by introducing the commutator
of two operators , and noting from the product rule that
(In the language of Lie groups and Lie algebras, this tells us that are (up to complex constants) the standard Lie algebra generators of the Heisenberg group.) From a quantum mechanical perspective, this lack of commutativity is the root cause of the uncertainty principle that prevents one from simultaneously localizing in both position and momentum past a certain point. Here is one basic way of formalising this principle:
Exercise 2 (Heisenberg uncertainty principle) For any
and
, show that
(Hint: evaluate the expression
in two different ways and apply the Cauchy-Schwarz inequality.) Informally, this exercise asserts that the spatial uncertainty
and the frequency uncertainty
of a function obey the Heisenberg uncertainty relation
.
Nevertheless, one still has the correspondence principle, which asserts that in certain regimes (which, with our choice of normalisations, corresponds to the high-frequency regime), quantum mechanics continues to behave like a commutative theory, and one can sometimes proceed as if the operators (and the various operators
constructed from them) commute up to “lower order” errors. This can be formalised using the pseudodifferential calculus, which we give below the fold, in which we restrict the symbol
to certain “symbol classes” of various orders (which then restricts
to be pseudodifferential operators of various orders), and obtains approximate identities such as
where the error between the left and right-hand sides is of “lower order” and can in fact enjoys a useful asymptotic expansion. As a first approximation to this calculus, one can think of functions as having some sort of “phase space portrait”
which somehow combines the physical space representation
with its Fourier representation
, and pseudodifferential operators
behave approximately like “phase space multiplier operators” in this representation in the sense that
Unfortunately the uncertainty principle (or the non-commutativity of and
) prevents us from making these approximations perfectly precise, and it is not always clear how to even define a phase space portrait
of a function
precisely (although there are certain popular candidates for such a portrait, such as the FBI transform (also known as the Gabor transform in signal processing literature), or the Wigner quasiprobability distribution, each of which have some advantages and disadvantages). Nevertheless even if the concept of a phase space portrait is somewhat fuzzy, it is of great conceptual benefit both within mathematics and outside of it. For instance, the musical score one assigns a piece of music can be viewed as a phase space portrait of the sound waves generated by that music.
To complement the pseudodifferential calculus we have the basic Calderón-Vaillancourt theorem, which asserts that pseudodifferential operators of order zero are Calderón-Zygmund operators and thus bounded on for
. The standard proof of this theorem is a classic application of one of the basic techniques in harmonic analysis, namely the exploitation of almost orthogonality; the proof we will give here will achieve this through the elegant device of the Cotlar-Stein lemma.
Pseudodifferential operators (especially when generalised to higher dimensions ) are a fundamental tool in the theory of linear PDE, as well as related fields such as semiclassical analysis, microlocal analysis, and geometric quantisation. There is an even wider class of operators that is also of interest, namely the Fourier integral operators, which roughly speaking not only approximately multiply the phase space portrait
of a function by some multiplier
, but also move the portrait around by a canonical transformation. However, the development of theory of these operators is beyond the scope of these notes; see for instance the texts of Hormander or Eskin.
This set of notes is only the briefest introduction to the theory of pseudodifferential operators. Many texts are available that cover the theory in more detail, for instance this text of Taylor.
— 1. Pseudodifferential operators —
The Kohn-Nirenberg quantisation was defined above for any symbol
obeying the very loose estimates (6). To obtain a clean theory it is convenient to focus attention to more restrictive classes of symbols. There are many such classes one can consider, but we shall only work with the classical symbol classes:
Definition 3 (Classical symbol class) Let
. A function
is said to be a (classical) symbol of order
if it is smooth and one has the derivative bounds
for all
and
. (Informally:
“behaves like”
, with each derivative in the frequency variable gaining an additional decay factor of
, but with each derivative in the spatial variable exhibiting no gain.) The collection of all symbols of order
will be denoted
. If
is a symbol of order
, the operator
is referred to as a pseudodifferential operator of order
.
As a major motivating example, any variable coefficient linear differential operator (3) of order will be a pseudodifferential operator of order
, so long as the coefficients
obey the bounds
for ,
, and
. (This would then exclude operators with unbounded coefficients, such as the harmonic oscillator, but can handle localised versions of these operators, and in any event there are other symbol classes in the literature that can be used to handle certain types of differential operators with unbounded coefficients.) Also, a fractional differential operator such as
will be a pseudodifferential operator of order
for any
. We refer the reader to Stein’s text for a discussion of more exotic symbol classes than the one given here.
The space of pseudodifferential operators of order form a vector space that is non-decreasing in
: any pseudodifferential operator of order
is automatically also of order
for any
. (Thus, strictly speaking, it would be more appropriate to say that
is a pseudodifferential operator of order at most
if
, but we will not adopt this convention for brevity.) The intuition to keep in mind is that a pseudodifferential operator of order
behaves like a variable coefficient linear differential operator of order
, with the obvious caveat that in the latter case
is restricted to be a natural number, whereas in the former
can be any real number. This intuition will be supported by the various components of the pseudodifferential calculus that we shall develop later, for instance we will show that the composition of a pseudodifferential operator of order
and a pseudodifferential operator of order
is a pseudodifferential operator of order
.
Before we set out this calculus, though, we give a fundamental estimate, which can be viewed as a variable coefficient version of the Hörmander-Mikhlin multiplier theorem:
Theorem 4 (Calderón-Vallaincourt theorem) Let
, and let
be a pseudodifferential operator of order
. Then one has
for all
. In particular,
extends to a bounded linear operator on each
space with
.
We now begin the proof of this theorem. The first step is a dyadic decomposition of Littlewood-Paley type. Let be a bump function supported on
that equals
on
. Then we can write
where
and
for . From dominated convergence, implies that
pointwise for . Thus by Fatou’s lemma, it will suffice to show that
uniformly in . Observe from Definition 3 and the Leibniz rule that each
is supported in the strip
and obeys the derivative estimates
for all .
From (5) and Fubini’s theorem we can express as an integral operator
for , where the integral kernel
is given by the formula
We can obtain several estimates on this kernel. Firstly, from the triangle inequality, (11), and the support property of we have the trivial bound
When , we may integrate by parts repeatedly, gaining factors of
at the cost of applying a
derivative to
for each such factor, and then if one applies the triangle inequality, (11), and support property of
as before we conclude that
for any ; by combining the estimates, we conclude that
for all and
. Differentiating (13) in
or
, and repeating the above arguments, we also obtain the estimates
Since the function has an
norm of
for any
, we now see from (12) and Young’s inequality that
Thus each component of
is under control (so for instance we may now discard the
term); the difficulty is to sum in
without losing any
-dependent factors. To do this, we first observe from (14), (15) and a routine summation of that the total kernel
(which is the integral kernel for
) obeys the pointwise bound
as well as the pointwise derivative bound
These are the usual kernel bounds for one-dimensional Calderón-Zygmund theory. From that theory we conclude that in order to prove the estimate (10), it suffices to establish the
case
From (17), we have already established a preliminary bound
for each , but a direct application of the triangle inequality will cost us a
-dependent factor, which we cannot afford. To do better, we need some “orthogonality” between the
. The intuition here is that each component
only interacts with the portion of
that corresponds to frequencies
of magnitude
, and that these regions are somehow “orthogonal” to each other. Informally, this suggests that
where is something like a Littlewood-Paley projection operator to frequencies
. If we accepted this heuristic, then we could informally use the Littlewood-Paley inequality (or
decoupling theory) to calculate
It is possible to make this approximation (19) more precise and establish (18): see Exercise 7. However, we will take the opportunity to showcase another elegant way to exploit “almost orthogonality”, known as the Cotlar-Stein lemma:
Lemma 5 (Cotlar-Stein lemma) Let
be bounded linear maps from one Hilbert space
to another
. Suppose that the maps
obey the operator norm bounds
for all
and some
, and similarly the maps
obey the operator norm bounds
for all
and some
. Then we have
Note that if the had pairwise orthogonal ranges then
would vanish whenever
, and similarly if the
had pairwise orthogonal coranges then the
would vanish whenever
. Thus the hypotheses of the Cotlar-Stein lemma are indeed some quantitative form of “almost orthogonality” of the
.
Proof: We use the method (which asserts that for a bounded linear map
between Hilbert spaces, the operator norm of
or
is the square of that of
or
). Applying this method to a single operator
we have
Taking geometric means we have
then by the triangle inequality we have . This loses a factor of
over the trivial bound. We can reduce this loss to
by a further application of the
method as follows. Writing
, we have
and similarly
so on taking geometric means we have .
We now reduce the loss in all the way to
by iterating the
method (this is an instance of a neat trick in analysis, namely the tensor power trick). For any integer
that is a power of two, we see from iterating the
method that
(In fact, this identity holds for any natural number , not just powers of two, as can be seen from spectral theory, but powers of two will suffice for the argument here.) We expand out the right-hand side and bound using the triangle inequality by
On the one hand, we can bound the norm by
grouping things slightly differently and using (22) twice, we can also bound this norm by
Taking the geometric mean, we can bound the norm by
Summing in using (20), then in
using (21) and so forth until the
sum (which is just summed with a loss of
), we conclude that
Sending , we obtain the claim.
Remark 6 There is a refinement of the Cotlar-Stein lemma for infinite series
of operators obeying the hypotheses of the lemma, in which it is shown that the series actually converges in the strong operator topology (though not necessarily in the operator norm topology); this refinement was first observed by Meyer, and can be found for instance in this note of Comech.
We will shortly establish the bounds
for any . The claim (18) then follows from the Cotlar-Stein lemma (using (17) to dispose of the
term).
We shall just show that
when ; the
case is treated similarly, as is the treatment of
(in fact this latter operator vanishes when
, though we will not really need this fact). We have
where
A direct application of (15) and the triangle inequality gives the bounds
(say), which when combined with Young’s inequality does not give the desired gain of . To recover this gain we begin integrating by parts. From (13) we have
Note that obeys similar estimates to
but with an additional gain of
. Thus the contribution of this term to
will be acceptable. The contribution of the other term, after an integration by parts, is
The kernel obeys the same bounds as (15) but with an additional gain of
; similarly from (16) the expression
obeys the same bounds as (15) but with an additional loss of
. The claim follows. This concludes the proof of the Calderón-Vaillancourt theorem.
Exercise 7 With the hypotheses as above, and with
a suitable Littlewood-Paley projection to frequencies
, establish the
operator norm bounds
for all
and
. Use this to provide an alternate proof of (18) that does not require the Cotlar-Stein lemma.
Now we give a preliminary composition estimate:
Theorem 8 (Preliminary composition) Let
be a pseudodifferential operator of some order
, and let
be a pseudodifferential operator of some order
. Then the composition
is a pseudodifferential operator of order
, thus there exists
such that
(note from Exercise 1 that
is uniquely determined).
Proof: We begin with some technical reductions in order to justify some later exchanges of integrals. We can express the symbol as a locally uniform limit of truncated symbols
as
, where
is a bump function equal to
near the origin; from the product rule we see that the symbol estimates (8) are obeyed by the
uniformly in
as long as
. If
is Schwartz, then so is
, and
can be verified to converge pointwise to
. If one can show that
for some pseudodifferential operator
of order
, with all the required symbol estimates (8) on
obeyed uniformly in
, then the claim will follow by using the Arzelà-Ascoli theorem to extract a locally uniformly convergent susbequence of the
and taking a limit. The upshot of this is that we may assume without loss of generality that the symbol
is compactly supported in
, so long as our estimates do not depend on the size of this compact support, but only on the constants in the symbol bounds (8) for
.
Similarly, we may approximate locally uniformly as the limit of symbols
that are compactly supprted in
, which makes
converge locally uniformly to
; from the compact support of
this also shows that
converges pointwise to
. From the same limiting argument as before, we may thus assume that
is compactly supported in
, so long as our estimates do not depend on the size of this support, but only on the constants in the symbol bounds (8) for
.
For , we have
hence on taking Fourier transforms
hence
and hence by Fubini’s theorem (and the compact support of and the Schwartz nature of
) we have
, where
for all and
, where the understanding is that the dependence of constants on
is only through the symbol bounds (8) for these symbols.
From differentiation under the integral sign and integration by parts we obtain the Leibniz identities
and
From this and an induction on (varying
as necessary, noting that if
maps
to
and
maps
to
) we see that to prove (25) it suffices to do so in the
case, thus we now only need to show that
for a given .
Applying a smooth partition of unity in the variable to
, it suffices to verify the claim in one of two cases:
-
is supported in the region
(so in particular
).
-
is supported in the region
.
(One can verify that applying the required cutoffs to do not significantly worsen the symbol estimates (8).) In the former case we write the left-hand side as
where
By repeating the proof of (14) we have
so from this and the symbol bound we obtain the claim in this case.
It remains to handle the latter case. Here we integrate by parts repeatedly in the variable to write the left-hand side of (26) as
for any . Then as before we can rewrite this as
where
By taking large enough we will eventually recover the bound
(in fact one can gain arbitrary powers of if desired), and so by repeating the previous arguments we also obtain the claim in this case.
The above proposition shows that if and
then
. The following exercise gives some refinements to this fact:
Exercise 9 (Composition of pseudodifferential operators) Let
and
for some
.
- (i) Show that
. (Hint: reduce as before to the case where
are compactly supported, and use the fundamental theorem of calculus to write
, where
. Then use the Fourier inversion formula, integration by parts, and arguments similar to those used to prove Theorem 8.
- (ii) Show that
, where
and
. (Hint: now apply the fundamental theorem of calculus once more to expand
.)
- (iii) Check (i) and (ii) directly in the classical case when
and
for some smooth
obeying the bounds (9) and for
. Based on this, for any integer
, make a prediction for an approximation to
as a polynomial combination of the symbols arbitrary
and finitely many of their derivatives which is accurate up to an error in
. Then verify this prediction.
Remark 10 From Exercise 9 we see that if
are pseudodifferential operators of order
respectively, then the commutator
differs from
by a pseudodifferential operator of order
, where
is the Poisson bracket
This approximate correspondence between the Lie bracket
(which plays a fundamental role in the dynamics of quantum mechanics) and the Poisson bracket
(which plays a fundamental role in the dynamics of classical mechanics) is one of the mathematical foundations of the correspondence principle relating quantum and classical mechanics, but we will not discuss this topic further here.
There is also a companion result regarding adjoints of pseudodifferential operators:
Exercise 11 (Adjoint of pseudodifferential operator) Let
.
- (i) If
is compactly supported, show that the function
defined by
is also a symbol of order
, and that
is the adjoint of
in the sense that
for all
.
- (ii) Show that even if
is not compactly supported, there is a unique pseudodifferential operator
of order
which is the adjoint of
in the sense that
for all
.
- (iii) Show that
is a pseudodifferential operator of order
.
Now we give some applications of the above pseudodifferential calculus.
Exercise 12 (Pseudodifferential operators and Sobolev spaces) For any
and
, define the Sobolev space
to be the completion of the Schwartz functions
with respect to the norm
- (i) If
is a non-negative integer, show that
for any
, thus in this case the Sobolev spaces agree (up to constants) with the classical Sobolev spaces (as discussed for instance in this set of notes).
- (ii) If
is a pseudodifferential operator of some order
, show that
for any
, thus
extends to a bounded linear map from
to
. (Hint: use Theorem 4 and Theorem 8).
- (iii) Let
be a pseudodifferential operator of some order
that obeys the strong ellipticity condition
for all
. Establish the Garding inequality
for all
and some
depending only on
. (Hint: use Exercises 9, 11 to express
as
for some pseudodifferential operators
of orders
and
respectively.) If
, deduce also the variant inequality
(possibly with slightly different choices of
).
The behaviour of pseudodifferential operators may be clarified by using a type of phase space transform, which we will call a Gabor-type transform.
Exercise 13 (Gabor-type transforms and pseudodifferential operators) Given any function
with the
normalisation
, and any
, define the Gabor-type transform
by the formula
thus
is the inner product of
with the function
, which is the “wave packet” formed from function
by translating by
and then modulating by
. (Intuitively,
measures the extent to which
lives at spatial location
and frequency location
.) We also define the adjoint map
for
by the formula
- (i) Show that for any
,
is a Schwartz function on
, thus
is a linear map from
to
. Similarly, show that for
,
is a Schwartz function on
, thus
is a linear map from
to
.
- (ii) Establish the identity
for any
, and conclude inparticular that
for any
, thus
extends to a linear isometry from
into
.
- (iii) For any smooth compactly supported
and
, establish the identity
where
is the (Kohn-Nirenberg) Wigner distribution of
, defined by the formula
and
is the phase space convolution
Remark 14 When
is a Gaussian, the transform
is essentially the Gabor transform (in signal processing) or the FBI transform (in microlocal analysis), and is also closely related to the Bargmann transform in complex analysis. There are some technical advantages with working with Gaussian choices of
, particularly with regards to the treatment of certain lower order terms in the pseudodifferential calculus; see for instance these notes of Tataru.
Note that is a Schwartz function on
, and by the Fourier inversion formula it has unit mass:
. (One also has the marginal distributions
and
, so
would be a strong candidate for a “phase space probability distribution” for
, save for the unfortunate fact that
has no reason to be non-negative. But even with oscillation,
still behaves like an approximation to the identity, so for
slowly varying
can be viewed as an approximation to
. Thus, Exercise 13(iii) can be intuitively viewed as saying that
behaves approximately like a multiplier in phase space:
Another informal way of viewing this assertion is that (for suitable choices of ) the translated and modulated functions
can be viewed as approximate eigenfunctions of
with eigenvalue
. This is for instance consistent with the approximate functional calculus
and
that one saw in Exercises 9, 11. The exercise below gives another way to view this approximation:
Exercise 15 (
![]()
bound) Let
be a smooth function obeying the “
bound”
for all
and
. Let
and
be as in Exercise 13. Show that there is a smooth kernel
obeying the bounds
for any
, such that
for any
. (Hint: work first in the case when
is compactly supported, where one can use Fubini’s theorem to derive an explicit integral expression for
, which one can then control by various integrations by parts.) Use this to establish the
bound
for any
; note that this gives an alternate proof of (18). (See also these notes of Tataru for further elaboration of this approach to pseudodifferential operators.)
As a sample application of the Gabor transform formalism we give a variant of the Garding inequality from Exercise 12(iii).
Theorem 16 (Sharp Garding inequality) Let
be a pseudodifferential operator of order
such that
for all
. Then one has
for all
, where
depends only on
.
Proof: From Exercise 11 we see that is a pseudodifferential operator of order
, hence by Exercise 12(ii) we have
Thus we may remove the imaginary part from and assume that
is real and non-negative. Applying a smooth partition of unity of Littlewood-Paley type, we can write
, where each
is also non-negative, supported on the region
, and obeys essentially the same symbol estimates as
uniformly in
. It then suffices to show that
uniformly in .
We now use the Gabor-type transforms from Exercise 13, except that we make
dependent on
. Specifically we pick a single real even
with
norm
, then define
for all
. We will approximate
by
Observe that
so by the triangle inequality it will suffice to establish the bound
However, it is not difficult (see exercise below) to show that is a symbol of order
uniformly in
, and the claim now follows from Exercise 12(ii).
Exercise 17 Verify the claim that
is a symbol of order
uniformly in
. (Here one will need the fact that
is a rescaling by a scaling factor
of
, which is an even Schwartz function of mean
. The even nature of
is needed to cancel some linear terms which would otherwise only allow one to obtain symbol bounds of order
rather than
.)
Remark 18 It is possible to improve the error term in the sharp Garding inequality, particularly if one uses the Weyl quantization rather than the Kohn-Nirenberg one (see Remark 19 below); also the non-negativity hypothesis on
can be relaxed in a manner consistent with the uncertainty principle; see this deep paper of Fefferman and Phong.
Remark 19 Throughout this set of notes we have used the Kohn-Nirenberg quantization
or equivalently (taking
to be compactly supported for sake of discussion)
However, this is not the only quantization that one could use. For instance, one could also use the adjoint Kohn-Nirenberg quantization
which one can easily relate to the Kohn-Nirenberg quantization by the identity
In particular, from Exercise 11 we see that if
is a symbol of order
, then
and
only differ by pseudodifferential operators of order
(and that both quantizations produce the same class of pseudodifferential operators of a given order). The operators
appearing earlier can also be viewed as a quantization of
(known as the anti-Wick quantization of
associated to the test function
). But perhaps the most popular quantization used in the literature is the Weyl quantization
which in some sense “splits the difference” between the Kohn-Nirenberg and adjoint Kohn-Nirenberg quantizations, being completely symmetric between the input spatial variable
and output spatial variable
. (Strictly speaking, this formula is only well-defined for say compactly supported symbols
; for more general symbols
one can define
in the weak sense as the distribution for which
for
(it is not difficult to use integration by parts to show that the expression in parentheses is rapidly decreasing in
, hence absolutely integrable). In particular there is now no error term in the analogue of Exercise 11:
All of the preceding theory for the Kohn-Nirenberg quantization can be adapted to the Weyl quantization with minor changes (for instance, the definition of the Wigner transform
changes slightly, and the operation
defined in (24) is replaced with the Moyal product), and as seen in Exercise 20 below, the two quantizations again produce the same classes of pseudodifferential operators, with symbols agreeing up to lower order terms.
Exercise 20 (Kohn-Nirenberg and Weyl quantizations are equivalent up to lower order) Let
be a real number.
- (i) If
is a symbol of order
, show that there exists a symbol
of order
such that
. Furthermore, show that
is a symbol of order
.
- (ii) If
is a symbol of order
, show that there exists a symbol
of order
such that
. Furthermore, show that
is a symbol of order
.
Exercise 21 (Comparison of quantizations) Let
be natural numbers, and let
be the monomial
.
- (i) Show that
.
- (ii) Show that
.
- (iii) Show that
, where
ranges over all tuples of operators consisting of
copies of
and
copies of
. For instance, if
, then
Informally, the Kohn-Nirenberg quantization always applies position operators to the left of momentum operators; the adjoint Kohn-Nirenberg quantization always applies position operators to the right of momentum operators; and the Weyl quantization averages equally over all possible orderings. (Taking formal generating functions, we also see (formally, at least) that the quantization of a plane wave
for real numbers
is equal to
in the Kohn-Nirenberg quantization,
in the adjoint Kohn-Nirenberg quantization, and
in the Weyl quantization.)
Exercise 22 (Gabor-type transforms and symmetries) Let
.
- (i) (Physical translation) If
and
is the function
, show that
for all
.
- (ii) (Frequency modulation) If
and
is the function
, show that
for all
.
- (iii) (Dilation) If
and
is the function
, show that
for all
, where
.
- (iv) (Fourier transform) If
, show that
.
- (v) (Quadratic phase modulation) If
and
is the function
, show that
for all
, where
.
We remark that the group generated by the transformations (i)-(v) is the (Weil representation of the) metaplectic group
.
Remark 23 Ignoring the changes in the Gabor test function
, as well as the various phases appearing on the right-hand side, we conclude from the above exercise that basic transformations on functions seem to correspond to various area-preserving maps of phase space; for instance, the Fourier transform is associated to the rotation
, which is consistent in particular with the fact that a fourfold iteration of the Fourier transform yields the identity operator. This is in fact a quite general phenomenon, with something asymptotically resembling such identities available for an important class of operators known as Fourier integral operators (but in higher dimensions one replaces the adjective with “area-preserving” with “symplectomorphism” or “canonical transformation“). However, as stated previously, the systematic development of the theory of Fourier integral operators is beyond the scope of this course.
Remark 24 Virtually all of the above theory extends to higher dimensions, and also to general smooth manifolds
as domains. In the latter case, the natural analogue of phase space is the cotangent bundle
, and the symplectic geometry of this bundle then plays a fundamental role in the theory (as already hinted at by the appearance of the Poisson bracket in Remark 10. See for instance this text of Folland for more discussion.
70 comments
Comments feed for this article
2 May, 2020 at 11:31 am
Tony Carbery
Hi Terry, I believe the observation about the strong operator topology in Remark 6 is due to Yves Meyer; in any case the argument is also sketched in Ch 7 Sect 5.3 of Stein’s Harmonic Analysis.
[Attribution added, thanks – T.]
2 May, 2020 at 12:58 pm
dn1214
In the syllabus of the course it is writen that the course would include paraproducts and the Walsh Carleson’s theorem. Are you still planning on giving lectures on these topics?
2 May, 2020 at 8:13 pm
Terence Tao
The next set of notes will cover Carleson’s theorem, though I have not yet decided whether to spend time on the Walsh version of the theorem. I had initially intended to also cover paraproducts, but a substantial amount of paradifferential calculus was already covered in the preceding 247A course, so I decided to replace that discussion with pseudodifferential calculus instead.
3 May, 2020 at 9:04 am
Anon
The two displays after (22) should have A^2, B^2 swapped?
[Corrected, thanks – T.]
3 May, 2020 at 9:32 am
Anon
In the final display of the proof of the Cotlar-Stein lemma, one seems to get
.
[I believe the estimate is correct as it stands – the estimate you claim is not dimensionally consistent (it isn’t invariant with respect to the operation of multiplying
by a scalar
). -T]
3 May, 2020 at 9:42 am
Anon
If this is true, then the argument still goes through assuming
(which can be assumed since the case
is trivial), since then
as
.
3 May, 2020 at 11:14 am
Anon
For
,
is dimensionally consistent.
3 May, 2020 at 7:58 pm
Terence Tao
Ah, I see now. I’ve adjusted the text accordingly (by using a different bound on the operator norms of the two tail terms one can avoid having to artificially introduce the ratio between A and B).
3 May, 2020 at 9:23 pm
Anon
:-)
3 May, 2020 at 11:01 am
Anon
After (17), it seems one wants
?
[Actually I prefer to work with the finite truncation here, but I have corrected a typo regarding this being the integral kernel for
(it should instead be
) -T]
4 May, 2020 at 3:35 am
Lior Silberman
While mathematicians tend to always write
for the semiclassical parameter, strictly speaking this is correct usage only for those who follow the physicists and define the Fourier transform by integration against
. For those who use the number theory normalization with
the correct semiclassical parameter is
(with
).
This is visible when you define the position operator with the factor
, which can clearly be multiplied by
to give the familiar
, but shouldn’t be multiplied by
.
4 May, 2020 at 7:29 pm
Anonymous
Dear Professor Tao,
What are the advantages/disadvantages of studying PDEs (Say Elliptic) using paraproducts and pseudodifferential operators, as oppose to using other stander methods like energy estimate, maximum principles, schauder estimates estimate. Thank you!
5 May, 2020 at 9:54 am
Terence Tao
Pseudodifferential operators are able to isolate specific regions of phase space; for instance it can be used in various wave equations to decompose a wave into “outgoing” and “incoming” waves (plus perhaps some error terms) which can be useful in various scattering theory type applications. They actually combine well with energy methods (for instance the “positive commutator method” is basically the energy method applied to pseudodifferential multipliers), and with various function space estimates such as Schauder estimates. Paraproducts are similarly able to isolate different types of frequency interactions (high-low, low-high, high-high), for instance leading to “microlocal gauge transformations” in which the most dangerous frequency interaction is isolated and tamed through an appropriate paradifferential gauge transformation, possibly at the cost of generating some uglier but more tractably estimated terms in the equation.
It is true though that such tools are somewhat non-local in nature and hence are not perfectly compatible with pointwise tools such as the maximum principle. (But in recent years non-local versions of the maximum principle have also been developed, so perhaps in the future one will see these tools cooperate more with each other.)
11 May, 2020 at 3:48 pm
Anonymous
I guess applying pseudodifferential operators to nonlinear equations does not work so well. What are some ways to get around this?
12 May, 2020 at 9:06 am
Terence Tao
The standard way to apply linear methods to nonlinear PDE is via perturbation theory; express the nonlinear PDE as a linear PDE plus a forcing term that depends nonlinearly on the solution, and try to show that the contribution of the nonlinear term is somehow controlled by the linear one using linear (or in some cases, bilinear, multilinear, or fully nonlinear) estimates. For the latter pseudodifferential operators can certainly be used. Of course, such techniques are only expected to be useful in perturbative regimes, where the solution or data is very close to an exact solution (e.g., if the data is small in some function space norm, so that the solution is expected to be close to the zero solution), or if one works only in a small region of space and time. However it can still be possible sometimes to extend such local results to global results, for instance there are certainly situations in which some energy-type quantity constructed using pseudo-differential operators enjoys some approximate monotonicity in time even after the nonlinearity is taken into account (this is particularly likely to happen if the nonlinearity is “defocusing” or “repulsive” in some sense), perhaps using variants of the Garding inequalities mentioned in this post.
6 May, 2020 at 9:29 pm
jair201p
Reblogged this on jair201p.
7 May, 2020 at 5:46 am
Charles Nash
The Fourier Transform F: Some simple observations that could form a student exercise.
F is unitary with respect to the usual inner product
F^4=I
Hence F has eigenvalues taken from the 4th roots of unity
All the roots are realised
The corresponding eigenfunctions hn say, are essentially
Hermite Hn polynomials times a Gaussian
The eigenspaces are infinite dimensional
and each such space corresponds to the value of n mod 4
i.e.
F(hn)=i^n hn
[A related exercise has now been added – T.]
7 May, 2020 at 1:13 pm
Anonymous
The projection operator on each of the 4 eigenspaces can be represented as a third degree polynomial of the Fourier transform.
8 May, 2020 at 10:40 pm
Charles Nash
Ah Thanks. Yes, if
is one of the eigenvalues, that polynomial seems to be

8 May, 2020 at 11:40 pm
Charles Nash
Oops I thinkI need another
in the denominator.
9 May, 2020 at 12:05 am
Charles Nash
arggh numerator
7 May, 2020 at 7:24 am
dn1214
I really enjoyed the proof of the Calderon-Vaillancourt, I feel like for the first time I understand it. Usually the textbook proof is the following: by integration by parthe the boundedness for operator of order m large enough is easy to obtain. Then one inducts on
to lower the degree of
, more precisely, if
is of order
then
is a sympbol of order
thus bounded and by duality, so is
. To go down to
, one takes
and
large enough for which the relevant operator to consider is
.
The two proofs seem really different, however they both rely on some
argument. Can anyone know a bit more on the difference of these two proofs, and would explain to me where ‘orthogonality’ is used in the latter?
7 May, 2020 at 5:41 pm
Terence Tao
Broadly speaking there are at least two general ways to establish inequalities
or estimates
. One way (what one might call the “divide and conquer” approach) is to split up
into pieces, apply various estimates and transforms to each of the pieces, and continue splitting up and estimating all the terms that arise until everything is satisfactorily controlled by
, and then sum up. The other is to try to express
, either exactly or approximately, as the sum of squares (or more generally as a sum or integral of manifestly non-negative quantities), either by using algebraic identities or some sort of monotonicity formula. Harmonic analysis arguments generally use the former type of argument, but the latter is more common in PDE, geometric analysis, and operator algebras. The latter approach is particularly good for proving sharp inequalities (e.g.,
with the optimal constant
), but it requires more algebraic structure on the objects being manipulated (and in some cases sum of squares decompositions are simply not available, cf. Hilbert’s seventeenth problem). Here, the algebraic structure is provided by the pseudodifferential calculus, which gives operator norm bounds on
that is basically
plus lower order terms. The divide and conquer approach in these notes (which also appears for instance in Stein’s “Harmonic analysis”) gives weaker bounds, but does not rely at all on the pseudodifferential calculus, which in this text I have arranged to come after the Calderon-Vaillancourt theorem is established rather than prior. (But note that the Garding and sharp Garding inequalities discussed in the post are proven using the sum-of-squares approach.)
[Edit: there are also many other ways to attack inequalities. For instance, one can take a variational approach and study how to maximise
or minimise
(or the difference or quotient of
and
) with respect to various parameters. Or one can try to transform the entire inequality
(rather than just the left-hand side or right-hand side) using tools such as duality or symmetry reduction. One can try to “categorify” an inequality into an injection (or surjection) in some combinatorial, geometric, probabilistic, or algebraic category. Qualitative forms of inequalities can sometimes be established by compactness arguments. One can sometimes exploit “gaps” (such as spectral gaps or integrality gaps) to obtain non-trivial inequalities as soon as some degenerate case is excluded. In some cases a strict inequality
can be established by a continuity argument, deforming both
and
to some degenerate case where the required inequality is easier to establish, and showing that one never passes through an equality case
in the course of this deformation. As far as I know, though, none of these strategies are particularly well-suited for establishing the Calderon-Vaillancourt theorem.]
8 May, 2020 at 11:21 am
Anonymous
It seems that there is some latex problem here.
[Can you be more specific? I do not see any issues. Note that the final paragraph is intended to be in italics. -T]
8 May, 2020 at 2:57 pm
Anonymous
Dear Prof. Tao:
Will you be teaching in the fall via zoom? Could you also make them accessible to the public?
9 May, 2020 at 9:10 am
Terence Tao
In the fall I will be teaching 246A (complex analysis), reusing the notes from https://terrytao.wordpress.com/category/teaching/246a-complex-analysis/ . UCLA has not yet decided what the format will be (it depends of course on the projected state of the pandemic in California by that time, which nobody knows with certainty), but it is likely some remote option will be offered even if the courses will resume in our physical classrooms.
8 May, 2020 at 3:59 pm
Lior Silberman
I was struggling with Exercise 1(i) all afternoon: the usual hypotheses are something like
, and the hypothesis (6) seemed too weak. The problem is that to gain a power
you would integrate (5) by parts
times, and this hits the symbol with a differentiation in
that under hypothesis (6) can lose us an arbitrary power of
and offset all our gains.
Eventually I found what (unless I’m missing something) is the obvious counterexample: take
. Then by (5) we have
for all Schwartz functions, and of course the constant function is not a Schwartz function.
8 May, 2020 at 4:01 pm
Lior Silberman
Sorry, I meant the usual hypothesis is
8 May, 2020 at 4:19 pm
Lior Silberman
With apologies for the comment barrage, assuming the symbol is Schwartz makes the operator continuous from tempered distributions to Schwartz functions.
I think the symbol class you want for Exercise 1(i) is
where
. The key point is that while differentiation with respect to
can lose arbitrary powers of
and differentiation with respect to
loses an arbitrary power of
the loss in
is less than the gain from the integration by parts.
9 May, 2020 at 9:08 am
Terence Tao
Huh, that is an unexpected subtlety that I had not realised! For these notes the
case will suffice, so that seems to be the simplest fix.
11 May, 2020 at 11:34 am
Anonymous
Shouldn’t = – due to integration by parts? for the momentum operator D.
11 May, 2020 at 12:04 pm
Anonymous
It’s correct as it stands because of the Hermitian inner product.
11 May, 2020 at 3:40 pm
Rex
In the proof of Lemma 5:
“Applying this method to a single operator {T_i} we have
\displaystyle \|T_i\|_{op} = \| T_i T_i^* \|_{op}^{1/2} \leq A
and similarly
\displaystyle \|T_i\|_{op} = \| T_i^* T_i \|_{op}^{1/2} \leq B. \ \ \ \ \ (22)
Taking geometric means we have…”
I think the A and B here should be switched (compare with the statement of Lemma 5)
[Corrected, thanks – T.]
11 May, 2020 at 8:47 pm
Anonymous
Is there a reason why multipliers are called “symbols”?
12 May, 2020 at 9:10 am
Terence Tao
I believe the term originates from the classical methods of symbolic calculus (also known as operational calculus) used to solve linear ODE by formally manipulating the polynomial that we would now call the symbol of the linear differential operator. I found for instance this historical article focusing in particular of the contributions of Heaviside (as well as Cauchy, Gregory, and Boole): https://www.jstor.org/stable/41133808
12 May, 2020 at 3:56 pm
Anonymous
Dear Professor Tao, just curious but are you of Shanghainese Wu speaking ancestry?
12 May, 2020 at 4:48 pm
Anonymous
What? How could one know? DNA test?
12 May, 2020 at 4:46 pm
Anonymous
How to show
? Here is my calculation, not sure what went wrong.



, which is not equation to
defined in equation (1).
12 May, 2020 at 5:07 pm
Terence Tao
You are using
to denote both a free variable and a bound variable; I recommend using something like
for the bound variable. Also, you dropped a factor of
after the third equals sign. At some point you will need to apply the Fourier inversion formula, as well as either an integration by parts or a differentiation under the integral sign. (You may find it easier to establish the equivalent identity
.)
12 May, 2020 at 6:05 pm
Anonymous
Hi, professor
I still don’t quite understand the intuition you mentioned in class: the symbol is about constant in rectangles in the phase plane with side length 2^k in the frequency variable and side length 1 in the physical variable. Can you explain what is the meaning of being about constant and how does this motivate the idea of applying Littlewood-Paley decomposition in the proof of Calderon-Vallaincourt theorem
12 May, 2020 at 7:18 pm
Terence Tao
Let’s say
is a symbol of order
, then
,
, and
when
. Thus, by the mean value theorem, we have
when
and
(at least if
have the same sign). The error term here is then lower order compared to the main term
when
and
, thus we expect
to behave like a constant on
rectangles in the region
. This partitions phase space into a collection of rectangles of various sizes, but the geometry of this partition becomes simpler if one restricts to a single annular region
in which one now just has a regularly spaced grid of rectangles, which is why a Littlewood-Paley decomposition would be a suggested tool of choice for analysing the pseudodifferential operator associated to this symbol.
13 May, 2020 at 5:36 am
Anonymous
Maybe a dumb question, but shouldn’t commutator and Poisson bracket differ by an operator of order
in Remark 10? I thought the argument there is that
plus an error which has the same order as in Exercise 9ii).
[Corrected, thanks – T.]
14 May, 2020 at 4:27 pm
Anonymous
I have two questions in the proof of theorem 8. In the second half, do we need to apply partition of unity to symbol a to get the desired estimates on kernel K_x as we already assume the symbol a to be compactly supported in both variables? I think i might miss something when i tried to build the estimate for the case \xi is close to \eta and \xi is away from \eta without applying a partition of unity to symbol a.
Also why is the case that \xi is close to \eta is the main contribution of the integral (26) so that we are not worried about the case \xi is away from \eta after we can have a good estimate on K_x for the case that \xi is close to \eta.
14 May, 2020 at 7:13 pm
Terence Tao
If one does not restrict the support of
to the region where
is close to
, then one does not get good bounds on the kernel
, for instance one cannot establish the bound
because the contribution of very large
(for large positive
) or very small
(for large negative
) becomes problematic.
The case when
is far away from
is expected to be negligible from stationary phase heuristics, because the phase
becomes non-stationary in the
variable in this regime.
15 May, 2020 at 12:43 pm
Anonymous
Hi, professor. In the last part of proof of theorem 8, when we consider the two cases where \eta is close to \xi and \eta is far from \xi, is there a special meaning to consider |\eta-\xi| >= 1/4 instead of |\eta-\xi| >= 1/2 ?
15 May, 2020 at 4:42 pm
Terence Tao
One can take other constants here than 1/2 and 1/4. But the two constants in the decomposition here need to be separated for each other in order to be able to use a smooth partition of unity to reduce to the two cases: if for instance one had the same constant (e.g., 1/2) in both cases then the only way to partition
into pieces with the indicated supports would be to use a rough cutoff such as
, which would destroy the symbol estimates as this cutoff is not differentiable.
15 May, 2020 at 2:00 pm
Anonymous
Dear professor Tao.
I think in the first lecture of the note 3 you mentioned that there is a reason why you place the theory of pseudodifferential operator here concerning come connection with the wave packet decomposition (maybe i misheard it). Can you briefly explain what is the connection and can you provide a reference for the wave packet decomposition (i don’t know anything about this decomposition)? Thank you.
15 May, 2020 at 4:49 pm
Terence Tao
The translated and modulated functions
that appear in the Gabor-type transforms are examples of wave packets (in this case, they are basically adapted to tiles
of unit length in phase space). In Notes 4 wave packets
adapted to other tiles are utilised.
16 May, 2020 at 9:36 am
Singular Integrals – Zeros and Ones
[…] 2) We assume(!) that is bounded from to , i.e., . This seems to a be a pretty strong assumption (we are assuming something that we really would like to prove). In practice to figure out when a given operator is bounded in usually is a simple task and almost follows from the Plancherel’s theorem (as we did with Hilbert transform, and similarly can be done for any convolution type operator as long as the Fourier transform of its kernel is in ). However, there are other operators when it is really a difficult task to prove boundedness even in (in this case there are other arguments: a) argument; b) Cotlar–Stein lemma (see for instance the proof of Calderón-Vallaincourt theorem, i.e., Theorem 4 in Tao’s notes)). […]
16 May, 2020 at 9:51 am
Fourier multipliers: examples on the torus – Zeros and Ones
[…] , i.e., our Fourier multipliers to depend on , namely These inequalities arise when studying pseudodifferential operators. Such multipliers create issues with orthogonality, namely the identity may not hold anymore if we […]
17 May, 2020 at 7:20 pm
Calvin Khor
I’m not sure how to use Exercise 7 to rigourise (18), which is the
bound uniform in
for
in the proof of the Calderon-Vallaincourt Theorem. When I simply replace
, I find that the exponential decay
is not enough to deal with the double sum in
: I get a bound linear in
. So I feel like I’m missing something quite simple. Anyone have some pointers?
17 May, 2020 at 7:37 pm
Terence Tao
Apply Littlewood-Paley decomposition to both sides of
(not just the left side) to get two useful exponential decay factors rather than just one.
18 May, 2020 at 3:25 am
Calvin Khor
Thanks for the hint. I managed to finish by instead using that
vanishes for
. I’m still not sure how to get a second exponential decay factor, because I only have one copy of
which can only use up one of the
s?
Also some minor typos
1. in exercise 8(i), the last few words, it should be Theorem 8, not Exercise 8
2. Near the beginning of the application of Cotlar-Stein, I think
should not have a star)
has one too many stars on the left (I believe
18 May, 2020 at 12:50 pm
Terence Tao
Thanks for the corrections.
The operator norm of
can be bounded both by
and by
, so by taking geometric means one can get a bound
that has two exponential decay factors (note that the precise numerical exponent in the exponential decay factors are not of major importance for these arguments). As you say one can certainly do better by exploiting the vanishing of
when
, although this is a feature specific to the Kohn-Nirenberg calculus and does not hold for instance for the Weyl calculus.
26 June, 2022 at 10:46 am
Connor
I’m trying to follow the argument Professor Tao has outlined but I end up with a dependence on
. To summarize, here is what I’ve done:
Applying Littlewood-Paley decompositions we have
Taking the geometric mean of the two inequalities we derived we have the estimate
.
It seems to me that Professor Tao is indicating that
is independent of
. My question is how can we show that explicitly? Summing the series by splitting it into the parts where
,
,
and
allows us to deal with the series in
and
but leaves us with a dependence on
. I think this is similar to the issue Calvin first encountered when using the single Littlewood-Paley decomposition.
If anyone has any clarifications or pointers I would be very appreciative!
30 June, 2022 at 6:22 am
Terence Tao
This triple sum is unbounded (consider the contribution of the diagonal terms
). To get operator norm bounds on
that are independent of
, one needs to exploit the almost orthogonality of the Littlewood-Paley projections, for instance via the Cotlar-Stein lemma.
6 July, 2022 at 7:18 am
Connor
Thank you, Professor Tao! My confusion was due to the fact that in the exercise you state we should be looking to prove (18) without using the Cotlar-Stein lemma.
6 July, 2022 at 9:50 am
Terence Tao
In this particular case one can argue without using Cotlar-Stein, using instead estimates such as
and
which can be easily proven by Plancherel. One can also take advantage of the approximate idempotency
where
is a suitable enlargement of
.
22 May, 2020 at 2:06 pm
247B, Notes 4: almost everywhere convergence of Fourier series | What's new
[…] this symbol is far too rough for us to be able to use pseudodifferential operator tools from the previous set of notes. Nevertheless, the “time-frequency analysis” mindset of trying to efficiently decompose […]
24 May, 2020 at 5:42 am
anon
I was just thinking about some earlier blog post you had made about problems with finding regular solutions to Navier-Stokes. I think the argument was that a potential problem was that the energy would increasingly be concentrated at smaller length-scales as time goes on. Just based upon physical intuition, if all energy would be concentrated at small length scales I would expect a lot of small eddies spinning in different directions. In that case, as the eddies get smaller I would expect the “surface area” between them to increase, which should also increase the viscous forces that should scale with the surface area of oppositely moving segments of the fluid. My guess is that the viscous forces should eventually quench the eddies at a sufficiently small scale. Just a thought!
26 May, 2020 at 9:08 pm
Anonymous
The proposed blowup mechanism doesn’t necessarily involve many concentrations close together, it should be possible with just one.
24 May, 2020 at 7:55 am
Xiaoyan Su
I can’t get uniform bounds in
for exercise 17. What I did was I expanded using Taylor’s with integral remainder
; the first term is exactly what we’re subtracting from and the gradient term disappears, so I just need to estimate the remaining integral correctly. But when I do this, I think I get the correct symbol estimate, but its constant in $k$ (hence bad when summed in $k$). Does this sound like its mostly correct, and do you have any hints?
26 May, 2020 at 1:45 pm
Terence Tao
This is the right approach (though when
is really large then the Taylor expansion is inefficient and other means of estimation may be superior). Intuitively
should be located in roughly the same location of phase space as
and in particular should exhibit decay when
is much larger than or much smaller than
, which can be used to recover summability in
.
26 May, 2020 at 5:45 pm
Xiaoyan Su
Thank you very much Prof Tao, I used the integral form for the remainder of the Taylor expansion, and I think my form of the remainder is good for large y, eta as well. I didn’t understand that intuition because
is very spread out in frequency for
, but with your hint I managed to finish the exercise.
27 May, 2020 at 1:06 pm
hhy177
In Exercise 20(i), I write Kohn-Nirenberg of a and Weyl quantization of
as integration against a kernel. The equality of the kernel gives me a formula for
as a necessary condition. To show it is sufficient, I need to use the kernel representation of the quantizations. Even if I assume a is compact support,
does not seem to be compact supported so I can only use the kernel representation for
but not for
. So it seems like I can only show
as tempered distribution for every Schwartz
. How do we (do we need to) show
as Schwartz function? (I guess we haven’t show the Weyl quantization of a symbol takes Schwartz to Schwartz, we did it only for KN quantization in Ex 1, but I think it is true as well).
27 May, 2020 at 2:59 pm
Terence Tao
For smooth compactly supported
, the associated Weyl symbol
will not be compactly supported in general, but one can still establish a lot of regularity and decay on
that will be sufficient to justify the formal calculations.
2 June, 2020 at 12:37 am
MATH 247B: Modern Real-Variable Harmonic Analysis – Countable Infinity
[…] Pseudifferential Operators […]
20 June, 2020 at 8:50 pm
Anonymous
man u ever take that semiclassical approximation of the WKB to the N to the I to the G-G-A, man?
i got my bro egorov helping me tame the singularities for the propagation this wave i’ve surfing brah
i’m only 12 so sorry
13 June, 2022 at 2:45 am
dn1214
Is there a generalization of the symbolic calculus for Lipschitz in
symbols
?
where we smoothly localize in
around
. (say that on the support of
,
). Is there a way of inverting the operator
in this way: there exists some operator
(of order
) such that
where
has lower order?
For example the elliptic symbol
The problem I can see is that the formula for commuting the principal symbol of
does not make sense anymore. But I have the intuition that the result I am asking for should be true.
18 June, 2022 at 7:58 am
Terence Tao
As a general principle, once one is outside the standard symbol classes, it becomes difficult to generate a useful symbolic calculus for general classes of these operators, and one should instead treat these operators with more specialized tools. For instance, inverting elliptic operators with Lipschitz (or even bounded) coefficients is a well studied topic in elliptic PDE, but requires tools specific to elliptic equations rather than general symbol calculus; see for instance the text by Gilbarg and Trudinger.