Note: this post is not required reading for this course, or for the sequel course in the winter quarter.
In a Notes 2, we reviewed the classical construction of Leray of global weak solutions to the Navier-Stokes equations. We did not quite follow Leray’s original proof, in that the notes relied more heavily on the machinery of Littlewood-Paley projections, which have become increasingly common tools in modern PDE. On the other hand, we did use the same “exploiting compactness to pass to weakly convergent subsequence” strategy that is the standard one in the PDE literature used to construct weak solutions.
As I discussed in a previous post, the manipulation of sequences and their limits is analogous to a “cheap” version of nonstandard analysis in which one uses the Fréchet filter rather than an ultrafilter to construct the nonstandard universe. (The manipulation of generalised functions of Columbeau-type can also be comfortably interpreted within this sort of cheap nonstandard analysis.) Augmenting the manipulation of sequences with the right to pass to subsequences whenever convenient is then analogous to a sort of “lazy” nonstandard analysis, in which the implied ultrafilter is never actually constructed as a “completed object“, but is instead lazily evaluated, in the sense that whenever membership of a given subsequence of the natural numbers in the ultrafilter needs to be determined, one either passes to that subsequence (thus placing it in the ultrafilter) or the complement of the sequence (placing it out of the ultrafilter). This process can be viewed as the initial portion of the transfinite induction that one usually uses to construct ultrafilters (as discussed using a voting metaphor in this post), except that there is generally no need in any given application to perform the induction for any uncountable ordinal (or indeed for most of the countable ordinals also).
On the other hand, it is also possible to work directly in the orthodox framework of nonstandard analysis when constructing weak solutions. This leads to an approach to the subject which is largely equivalent to the usual subsequence-based approach, though there are some minor technical differences (for instance, the subsequence approach occasionally requires one to work with separable function spaces, whereas in the ultrafilter approach the reliance on separability is largely eliminated, particularly if one imposes a strong notion of saturation on the nonstandard universe). The subject acquires a more “algebraic” flavour, as the quintessential analysis operation of taking a limit is replaced with the “standard part” operation, which is an algebra homomorphism. The notion of a sequence is replaced by the distinction between standard and nonstandard objects, and the need to pass to subsequences disappears entirely. Also, the distinction between “bounded sequences” and “convergent sequences” is largely eradicated, particularly when the space that the sequences ranged in enjoys some compactness properties on bounded sets. Also, in this framework, the notorious non-uniqueness features of weak solutions can be “blamed” on the non-uniqueness of the nonstandard extension of the standard universe (as well as on the multiple possible ways to construct nonstandard mollifications of the original standard PDE). However, many of these changes are largely cosmetic; switching from a subsequence-based theory to a nonstandard analysis-based theory does not seem to bring one significantly closer for instance to the global regularity problem for Navier-Stokes, but it could have been an alternate path for the historical development and presentation of the subject.
In any case, I would like to present below the fold this nonstandard analysis perspective, quickly translating the relevant components of real analysis, functional analysis, and distributional theory that we need to this perspective, and then use it to re-prove Leray’s theorem on existence of global weak solutions to Navier-Stokes.
— 1. Quick review of nonstandard analysis —
In this section we quickly review the aspects of nonstandard analysis that we need. Let denote the “standard” universe of “standard” mathematical objects; this includes what one might think of as “primitive” standard objects such as (standard) numbers and (standard) points, but also sets of standard objects (such as the set
of real numbers, or the Euclidean space
), or functions
from one standard space to another, or function spaces such as
of such functions (possibly quotiented out by almost everywhere equivalence), and so forth. In short,
should contain all the standard objects that one generally works with in analysis. One can require that this universe obey various axioms (e.g. the Zermelo-Fraenkel-Choice axioms of set theory), but we will not be particularly concerned with the precise properties of this universe (we won’t even need to know whether
is a set or a proper class).
What nonstandard analysis does is take this standard universe of standard objects and embed it in a larger nonstandard universe
of nonstandard objects which has similar properties to the standard one, but also some additional properties. As discussed in this previous post, the relationship between the standard universe
and the nonstandard universe
is somewhat analogous to that between the rationals
and its metric completion
; most of the algebraic properties of
carry over to
, but
also has some additional completeness and (local) compactness properties that
lacks. Also, one should think of
as being far “larger” than
, in much the same way that
is larger than
in various senses, for instance in the sense of cardinality.
There is one important subtlety concerning the nonstandard universe : it comes with a more restrictive notion of subset (or of function) than the “external” notion of subset or function that one has if one views
from some external metatheory (e.g., if one places both
and
inside a very large model of ZFC). Thus, for instance, an externally defined subset of the nonstandard reals
may or may not be an internal subset of these reals (in particular, the embedded copy of the standard reals
is not an internal subset of
, being merely an external subset instead); similarly, an externally defined function from
to
need not be an internal function (for instance, the standard part function
will be external rather than internal). The relationship between internal sets/functions and external sets/functions in nonstandard analysis is somewhat analogous to the relationship between measurable sets/functions and arbitrary sets/functions in measure theory.
The reals can be constructed from the rationals
in a number of ways, such as by forming Cauchy sequences in
and quotienting out by the sequences that converge to zero; similarly, the nonstandard universe
can be formed from the standard one
in a number of ways, such as by forming arbitrary sequences in
and quotienting out by a non-principal ultrafiter. See for instance this previous post for details. However, much as the precise construction of the reals
is often of little direct importance in applications, we will not need to care too much about how the nonstandard universe is constructed. Rather, the following properties of this universe will be used:
- (i) (Embedding) Every standard object, space, operation, or function
in
has a nonstandard counterpart
in
. For instance, if
is a real number in the set
of standard reals, then
will be an element of the set
of nonstandard reals; if
is a standard function, then
is a nonstandard function from the nonstandard Euclidean space
to the nonstandard reals
. The standard addition operation
on the standard reals
induces a nonstandard addition operation
on the nonstandard reals, though to avoid notational clutter we will write
as
, and similarly for other basic mathematical operations. Similarly, the norm function
has a nonstandard counterpart
that assigns a nonstandard non-negative real
to any nonstandard
function
. (To avoid notational clutter, we will often abuse notation by identifying
with
for various “primitive” mathematical objects
such as real numbers, arithmetic operations such as
, or functions such as
, unless we have a pressing need to carefully distinguish a standard object
from its representative
in the nonstandard universe.)
- (ii) (Transfer) If
is a standard predicate in first order logic involving some finite number of standard objects
(with
a fixed standard natural number), and possibly some quantification over standard sets, and
is the nonstandard version of the predicate in which one quantifies over nonstandard sets, then
is true if and only if
is true. Important caveat: the predicate
needs to be internal to the mathematical language used internally to both
and
separately; it is not allowed to use external concepts dependent on the way in which
embeds into
, or how either universe embeds into an external metatheory.
- (iii) (
–saturation) Let
be standard natural numbers, and suppose that for each standard natural number
,
is a nonstandard predicate on
nonstandard variables
and nonstandard constants
. If any finite collection of the predicates
are simultaneously satisfiable (thus, for each standard
, there exist nonstandard objects
such that
holds for all
), then the entire collection
is simultaneously satisfiable (thus there exists nonstandard objects
such that
holds for all
).
The -saturation property (also informally referred to as countable saturation, though this is technically a slight misnomer) resembles the finite intersection property that characterises compactness of topological spaces (and can thus be viewed as somewhat analogous to the local compactness property for the reals
), except that the finite intersection property involves arbitrary families of (closed) sets, whereas the
-saturation property requires the collection of predicates involved to be countable. It is possible to construct nonstandard models with a higher degree of saturation (where one can use more predicates
, as long as the total number does not exceed some cardinal
which relates to the size of the nonstandard universe
), for instance by replacing the sequences used to construct the nonstandard universe with tuples ranging over a larger cardinality set. This may potentially be useful for certain types of analysis, for instance ones involving non-separable spaces, or Frechet spaces involving an uncountable number of seminorms.
Let us take for granted the existence of a nonstandard universe obeying the embedding, transfer, and saturation properties, and see what we can do with them. Firstly, transfer shows that the map is injective:
if and only if
. The field axioms of the standard reals
can be phrased in the language of first-order logic, and hence by transfer the nonstandard reals
also form a field. For instance, the assertion “For every non-zero standard real
, there exists a standard real
such that
” transfers over to “For every non-zero nonstandard real
, there exists a nonstandard real
such that
“. If
is a standard natural numer, one can transfer the statement “
if and only if
” from standard tuples to nonstandard tuples; among other things, this gives the nice identification
when
is a standard natural number. (The situation is more subtle when
is a nonstandard natural number, but in most PDE applications one works in a fixed dimension
and will not need to deal with this subtlety.) As one final example, “If
, then
holds if and only if
” transfers to “If
, then
holds if and only if
“. More generally, basic inequalities such as Hölder’s inequality, Sobolev embedding, or the Bernstein inequalities transfer over to the nonstandard setting without difficulty.
As a basic example of saturation, for each standard natural number let
denote the statement “There exists a nonstandard real
such that
“. These statements are finitely satisfable, hence by
-saturation they are jointly satisfiable, thus there exists a nonstandard real
which is unbounded in the sense that it is larger than every standard natural number (and hence also by every standard real number, by the Archimedean property of the reals). Similarly, there exist nonstandard real numbers
which are non-zero but still infinitesimal in the sense that
for every standard real
.
On the other hand, one cannot apply the saturation property to the statements “There exists a nonstandard real such that
and
“, since
is not known to be an internal subset of the nonstandard universe
and so cannot be used as a constant for the purposes of saturation. (Indeed, since this sequence of statements is finitely satisfiable but not jointly satisfiable, this is a proof that
is not an internal subset of
, and must instead be viewed only as an external subset.)
Now we develop analogues of the sequential-based theory of limits in nonstandard analysis. The following dictionary may be helpful to keep in mind when comparing the two:
Standard real |
A real number |
Nonstandard reals |
A sequence |
Embedding |
A constant sequence |
Internal set |
A sequence |
Embedding |
A constant sequence |
External set |
A collection of sequences of reals |
Internal function |
A sequence |
Embedding |
A constant sequence |
External function |
A map from sequences of vectors to sequences of reals |
Equality |
After passing to a subsequence, |
|
|
|
|
|
Bounded sequences |
|
Sequences converging to zero (possibly after passing to subsequence) |
|
Convergent sequences (possibly after passing to subsequence) |
|
Bolzano-Weierstrass theorem |
Standard part |
Limit |
Note in particular that in the nonstandard analysis formalism there is no need to repeatedly pass to subsequences, as is often the case in sequential-based analysis.
A nonstandard real is said to be bounded if one has
for some standard
. In this case, we write
, and let
denote the set of all bounded reals. It is an external subring of
that in turn contains
as a external subring.
A nonstandard real is said to be infinitesimal if one has
for all standard
. In this case, we write
, and let
denote the set of all infinitesimal reals. This is another external subring (in fact, an ideal) of
, and
can be viewed as external vector spaces over
.
The Bolzano-Weierstrass theorem is fundamental to orthodox real analysis. Its counterpart in nonstandard analysis is
Theorem 1 (Nonstandard version of Bolzano-Weierstrass) As external vector spaces over
, we have the decomposition
.
Proof: The only real which is simultaneously standard and infinitesimal is zero, so . It thus suffices to show that every bounded real
can be written in the form
for some standard
. But the set
is a Dedekind cut; setting
to be the supremum of this cut, we have
for all standard natural numbers
, hence
as desired.
If and
for some standard real
, we call
the standard part of
and denote it by
: thus
is the linear projection from
to
with kernel
. It is an algebra homomorphism (this is the analogue of the usual limit laws in real analysis).
In real analysis, we know that continuous functions on a compact set that are pointwise bounded are automatically uniformly bounded. There is a handy analogue of this fact in nonstandard analysis:
Lemma 2 (Pointwise bounded/infinitesimal internal functions are uniformly bounded/infinitesimal) Let
be an internal function.
- (i) If
for all
, then there is a standard
such that
for all
.
- (ii) If
for all
, then there is an infinitesimal
such that
for all
.
Proof: Suppose (i) were not the case, then the predicates “ and
” would be finitely satisfiable, hence jointly satisfiable by
-saturation. But then there would exist
such that
for all
, contradicting the hypothesis that
.
For (ii), observe that the predicates “ is a nonstandard real such that
for all
” are finitely satisfiable, hence jointly satisfiable by
-saturation, giving the claim.
Because the rationals are dense in the reals, we see (from saturation) that every standard real number can be expressed as the standard part of a bounded rational, thus . This can in fact be viewed as a way to construct the reals; it is a minor variant of the standard construction of the reals as the space of Cauchy sequences of rationals, quotiented out by Cauchy equivalence.
Closely related to Lemma 2 is the overspill (or underspill) principle:
Lemma 3 Let
be an internal predicate of a nonnegative nonstandard real number
.
- (i) (Overspill) If
is true for arbitrarily large standard
, then it is also true for at least one unbounded
.
- (ii) (Underspill) If
is true for arbitrarily small standard
, then it is also true for at least one infinitesimal
.
Proof: To prove (i), observe that the predicates “ and
” for
a standard natural number are finitely satisfiable, hence jointly satisfiable by
-saturation, and the claim follows. The claim (ii) is proven similarly, using
instead of
.
Corollary 4 Let
be an internal function. If
for all standard
, then one has
for at least one unbounded
.
Proof: Apply Lemma 3 to the predicate .
The overspill principle and its analogues correspond, roughly speaking, to the “diagonalisation” arguments that are common in sequential analysis, for instance in the proof of the Arzelá-Ascoli theorem.
— 2. Some nonstandard functional analysis —
In the discussion of the previous section, the real numbers could be replaced by the complex numbers
or finite-dimensional vector spaces
(with
a standard natural number) with essentially no change in theory. However the situation becomes a bit more subtle when one works with infinite dimensional spaces, such as the functional spaces that are commonly used in PDE.
Let be a standard normed vector space with norm
, then we can form the nonstandard function space
with a nonstandard (or internal) norm
. This is not quite a normed vector space when viewed externally, because the nonstandard norm
takes values in the nonstandard nonnegative reals to
rather than the standard nonnegative reals
. However, we can form the subspace
of
consisting of those vectors
which are strongly bounded in the sense that
. This is an external real subspace of
that contains
. It comes with a seminorm
defined by
It is easy to see that this is a seminorm. The null space of this seminorm is the subspace of
consisting of those vectors
which are strongly infinitesimal (in
) in the sense that
; we say two elements of
are strongly equivalent (in
) if their difference is strongly infinitesimal. In infinite dimensions,
is no longer locally compact, and the Bolzano-Weierstrass theorem now only gives an inclusion:
In general we expect to be significantly larger than
(this is the nonstandard analogue of the sequential analysis phenomenon that most bounded sequences in
will fail to have convergent subsequences). For instance if
is a standard Hilbert space with an orthonormal system
, and
is an unbounded natural number, one can check that
lies in
but is not in
. The quotient space
is a normed vector space that contains
as an isometric subspace and is known as the nonstandard hull of
, but we will not explicitly use this space much in these notes.
In functional analysis one often has an embedding of standard function spaces
, with an inequality of the form
for all
and some constant
. For instance, one has the Sobolev embeddings
whenever
,
, and
. One easily sees that such an embedding
induces also embeddings
and
.
If the embedding is compact – so that bounded subsets in
are precompact in
– then we can partially recover the missing inclusion in the Bolzano-Weierstrass theorem. This follows from
Lemma 5 (Compactness) Let
be a standard compact subset (in the strong topology) of a standard normed vector space
. Then one has the inclusion
that is to say every
can be decomposed (uniquely) as
with
and
.
This is the nonstandard analysis analogue of the assertion that compact subsets of a normed metric space are sequentially compact.
Proof: Uniqueness is clear (since non-zero standard elements of have non-zero standard, hence non-infinitesimal, nonstandard norm), so we turn to existence. If this failed for some
, then for every
, there exists a standard
such that
. Hence, by compactness of
, one can find a standard natural number
and standard
such that for all
, one has
for some
. By transfer, (viewing
as constants), this implies that for all
, one has
for some
. Applying this with
, we obtain a contradiction.
We also note an easy converse inclusion: if is a standard open subset of a standard normed vector space
, then
Exercise 6 Suppose that the standard normed vector space
is separable. Establish the converse implications, that a standard subset
of
is compact whenever
, and a standard subset
is open whenever
. (The hypothesis of separability can be relaxed if one imposes stronger saturation properties on the nonstandard universe than
-saturation.)
Exercise 7 Let
be a standard function on a standard subset
of a normed vector space
, and let
be the nonstandard counterpart.
- (i) Show that
is bounded if and only if
for all
.
- (ii) Show that
is continuous (in the strong topology) if and only if
whenever
,
are such that
is strongly equivalent to
.
- (iii) Show that
is uniformly continuous (in the strong topology) if and only if
whenever
are such that
is strongly equivalent to
.
- (iv) If
is compact, show that
. Conclude the well-known fact that a standard continuous function on a compact set
is uniformly continuous and bounded.
Now we have
Theorem 8 (Compact embeddings) If
is a compact embedding of standard normed vector spaces
, then
If furthermore the closed unit ball of
is compact (not just precompact) in
, we can sharpen this to
This is the analogue of Proposition 3 of Notes 2.
Proof: Since , it suffices to show that every
can be written as
with
and
. But this is immediate from Lemma 5 (applied to the closure in
of a closed unit ball in
).
From the above theorem and the Arzelá-Ascoli theorem, we see for instance that a nonstandard Lipschitz function from to
with bounded nonstandard Lipschitz norm can be expressed as the sum of a standard Lipschitz function and a function which is infinitesimal in the uniform norm. Here is a more general version of this latter assertion:
Exercise 9 (Nonstandard version of Arzelá-Ascoli) Let
be a standard open set, and let
be a nonstandard function obeying the following axioms:
- (i) (Pointwise boundedness) For all
, we have
.
- (ii) (Pointwise equicontinuity) For all
, we have
whenever
and
.
Let
denote the function
(this is well-defined by pointwise boundedness). Then
is continuous, and its nonstandard representative
is locally infinitesimally close to
in the sense that
whenever
. Conclude in particular that for every standard compact
there exists an infinitesimal
such that
for all
.
We remark that the above discussion for normed vector spaces also extends without difficulty to Frechet spaces that have at most countably many seminorms
, with
now consisting of those
with
for all
, and
now consisting of those
with
for all
.
Now suppose we work with the dual space of a normed vector space
. (Here unfortunately we have a clash of notation, as the asterisk will now be used both to denote nonstandard representative and dual; hopefully the mathematics will still be unambiguous.) A nonstandard element
of
(thus, a nonstandardly continuous linear functional
) is said to be weak*-bounded if
for all
, and weak*-infinitesimal if
. The space of weak*-bounded elements of
will be denoted
, and the space of weak*-infinitesimal elements denoted
. These are related to the strong counterparts
of these spaces by the inclusions
we also have the inclusion
For instance, if is a standard Hilbert space with orthonormal system
, and
is an unbounded natural number, then
lies in
(and hence also in
) and in
but not in
. For a more subtle example, if we form the nonstandard universe
by taking an ultrapower with respect to an ultrafilter that deems all sets of density zero to be negligible, and one takes
to be the ultralimit of the sequence
, then
also lies in
(and hence also in
) but not in
or
, basically because for any
, the inner products
converge to zero in density. (Thus we see a slight difference here between the classical notion of weakly bounded sequences, which on Banach spaces are equivalent to bounded sequences by the Banach-Steinhaus theorem.)
(One could also develop similar notations in which one uses weak topologies instead of the weak* topology, but we will not need the weak topology in these notes.)
We have the following nonstandard version of the Banach-Alaoglu theorem:
Theorem 10 (Nonstandard version of Banach-Alaoglu) If
is a normed vector space with dual
, then we have the inclusion
Proof: Since , it suffices to show that every
can be decomposed as
for some
and
.
Let , thus there is a standard
such that
for all
. In particular, if we define
for
, then
depends linearly on
and
for all
; thus
is an element of
. Setting
, we see from construction that
for all
, giving the claim.
Theorem 10 can be compared with Proposition 2 of Notes 2, except in this nonstandard analysis setting, no separability is required on the predual space .
If is a standard bounded linear operator between normed vector spaces, then one has a nonstandard linear operator
. It is easy to see that this operator maps
to
and
to
. The adjoint operator
similarly maps
to
and
to
, but also takes
to
and
to
.
A (nonstandard) linear operator will be said to be an approximate identity on
if
maps
to
. Here are two basic and useful examples of such approximate identities:
Exercise 11 (Frequency and spatial localisation as approximate identities)
- (i) For standard
and unbounded
, show that the nonstandard Littlewood-Paley projection
is an approximate identity on
, and more generally on the Sobolev spaces
for any standard natural number
.
- (ii) For standard
and unbounded
, and a standard test function
that equals
near the origin, show that the nonstandard spatial truncation opertor
is an approximate identity on
, and more generally on the Sobolev spaces
for any standard natural number
. What happens at
?
— 3. Nonstandard analysis and distributions —
Let be a standard open set. The dual of the standard space
of test functions on
is the standard space
of distributions. Like any other standard space, it has a nonstandard counterpart
, whose elements are the nonstandard distributions.
A nonstandard distribution will be said to be weakly bounded if, for any standard compact set
, there is a standard natural number
and standard
such that one has the bound
for all standard . (It would be slightly more accurate to use the terminology “weak-* bounded” instead of “weakly bounded”, but we will omit the asterisk here to make the notation a little less clunky. Similarly for the related concepts below) Thus for instance any standard distribution is weakly bounded, and if
is any normed vector space structure on
that is continuous in the test function topology, and
, then
will be weakly bounded. We say that a nonstandard distribution
is weakly infinitesimal if one has
for all standard
. For instance, if
is any normed vector space structure on
and
, then
will be weakly infinitesimal. We say that two nonstandard distributions
are weakly equivalent, and write
, if they differ by a weakly infinitesimal distribution, thus
for all standard test functions .
If is a weakly bounded distribution, we can define the weakly standard part
to be the distribution defined by the formula
This is the unique standard distribution that is weakly equivalent to . As such, it must be compatible with the other decompositions of the preceding section. For instance, if
is a normed vector space structure on
and
, then the decomposition
must agree with the one in Theorem 10, thus
and
. In particular, if
embeds compactly into a normed vector space
, we also have
, thus weak equivalence in
implies strong equivalence in
. By the definition of the dual norm we also conclude the Fatou-type inequality
whenever .
Informally, represents the portion of
that one can “observe” at standard physical scales and standard frequency scales, ignoring all components of
that are at unbounded or infinitesimal physical or frequency scales. The following examples may help illustrate this point:
Example 12 Let
be an unbounded natural number, and on
let
be the nonstandard distribution
. Then
is weakly infinitesimal, so
. This is in contrast to the pointwise standard part
, which is a rather wild (and almost certainly non-measurable) function from
to
. The nonstandard distribution
is not even pointwise bounded in general, so does not have a pointwise standard part, but is still weakly infinitesimal and so
. Similarly if one replaces
by
for some standard bump function
.
The following table gives the analogy between these nonstandard analysis concepts, and the more familiar ones from sequential weak compactness theory:
Standard distribution |
Distribution |
Nonstandard distribution |
Sequence of distributions |
Embedding |
Constant sequence |
|
|
|
|
|
|
|
|
|
|
|
|
|
Weak limit of |
Exercise 13 Let
, establish the Pythagorean identity
This identity can be used as a starting point for the theory of concentration compactness, as discussed in these notes. What happens in other
spaces?
Exercise 14 Show that every standard distribution
is the weakly standard part of some weakly bounded nonstandard test function
. (Hint: When
is
, one can convolve with a nonstandard approximation to the identity and also apply a nonstandard spatial cutoff. When
is a proper subset of
one also has to smoothly cut off outside an infinitesimal neighbourhood of the boundary of
if one wants to make the convolution well defined.) This result can be viewed as analogous to the previous observation that every standard real is the standard part of a bounded rational. This also provides an alternate way to construct distributions, as weakly bounded nonstandard test functions up to weak equivalence.
Exercise 15 (Arzela-Ascoli, nonstandard distributional form) If
is a nonstandard continuous function obeying the pointwise boundedness and pointwise equicontinuity axioms from Exercise 9. Show that
is also a weakly bounded distribution and that the weakly standard part of
agrees with the pointwise standard part:
. In particular, if
is also weakly infinitesimal, conclude that
for every standard compact set
.
— 4. Leray-Hopf solutions to Navier-Stokes —
For that is divergence-free, recall that a Leray-Hopf solution to the Navier-Stokes equations on
is a distribution
vanishing outside of
that has the additional regularity
for almost all , and solves the equation
in the sense of distributions. We now give a nonstandard interpretation of this concept:
Proposition 16 (Nonstandard interpretation of Leray-Hopf solution) Let
be divergence-free. Let
be a nonstandard smooth function obeying the following properties:
- (i) (Energy inequality) For all nonstandard
, one has the nonstandard energy inequality
Here of course we use the nonstandard version of the Lebesgue integral, and extend
to
in the usual fashion (i.e., we identify
with
).
- (ii) (Initial data) We have
.
- (iii) (Weak time regularity) For any standard
,
.
- (iv) (Navier-Stokes) On
, one has the nonstandard (classical) forced Navier-Stokes equation
where the forcing term
, after extending by zero to
, is weakly infinitesimal.
Then after extending
by zero to
,
is weakly bounded and
is a Leray-Hopf solution to Navier-Stokes. Conversely, every such Leray-Hopf solution arises in this fashion.
Thus, roughly speaking, Leray-Hopf solutions arise from nonstandard strong solutions to Navier-Stokes in which one permits weakly infinitesimal changes to the initial data and forcing term, which are arbitrary save for the constraint that these changes do not add more than an infinitesimal amount of energy to the system, and which also obey a technical but weak condition on the time derivative. Thus, for instance, one can insert a nonstandard frequency mollification at an unbounded frequency cutoff , or a spatial truncation at an unbounded spatial scale
, without difficulty (so long as one checks that such modifications do not introduce more than an infinitesimal amount of energy into the system), which can be used to recover the standard construction of Leray-Hopf solutions.
Proof: First suppose that obeys all the axioms (i)-(iv). We now repeat the arguments used to prove Theorem 14 of Notes 2, but translated to the language of nonstandard analysis. (All the key inputs are still basically the same; the changes are almost all entirely in the surrounding formalism.)
From (i) we see that
and
which implies that
and
Also, for every standard we have
and hence by Sobolev embedding
for sufficiently close to
. From Hölder this gives
and then by the boundedness of the Leray projector
We have
in the sense of nonstandard distributions. Taking weakly standard part, we will obtain a weak solution to the Navier-Stokes equations as long as
Both sides lie in , but the equality requires some care. Applying a standard test function
, it suffices to show that
for all such , which we can assume to be supported in
. One can check that
lies in
. If we can show that for every standard compact set
that
hence by Corollary 4 the same claim is true for some unbounded ; however, the nonstandard
norm of
outside of an unbounded ball
is infinitesimal, and we then conclude (6).
Fix . By Hölder’s inequality, it now suffices to show that
that is to say is strongly equivalent to
in
. By another application of Hölder, it suffices to show that
is strongly equivalent to
in
.
For unbounded ,
is strongly equivalent to
in
; since
is bounded in
, we also see from Bernstein’s theorem and the triangle inequality that
is strongly equivalent to
in
. Thus it suffices to show that for at least one unbounded
, that
is strongly equivalent to
in
. By overspill, it suffices to do this for arbitarily large standard
. But by Bernstein’s inequality, the difference
is bounded in
, has space derivative bounded in
, and time derivative bounded in
on
by hypothesis, hence is equicontinuous from the fundamental theorem of calculus and Cauchy-Schwarz; by Exercise 15 we conclude that
is strongly infinitesimal in
, and hence also in
as required.
To show the energy inequality for , we again repeat the arguments from Notes 2. If
is standard and
is a standard non-negative test function supported on
of total mass one for some small standard
, then from averaging (4) we have
Taking weakly standard parts using (1) we conclude that
The energy inequality (2) then follows from the Lebesgue differentiation theorem (which, incidentally, can also be translated into a nonstandard analysis framework, as discussed to some extent in this previous post).
Now we establish the converse direction. Let be a Leray-Hopf solution, then from Sobolev and Hölder we have
for some sufficiently close to
, and hence from the weak Navier-Stokes equation and Bernstein’s inequality we have
for every standard (but without claiming the bound to be uniform in
).
Now let be infinitesimal and
be unbounded, let
be a standard non-negative test function of total mass
supported on
, and let
be the nonstandard function
arising from performing a frequency truncation in space and a smooth averaging in time. This is a nonstandard smooth function. From Minkowski’s inequality, (2), and the non-expansive nature of on
we see that we have the energy inequality (4) (in fact we do not even need the
error here). From Exercise 19 of Notes 2, we know that
converges strongly in
to
as
, and hence
is strongly equivalent to
in
for all
, and hence
is also. From (7) and Minkowski’s inequality we also conclude property (iii) of the proposition.
The only remaining thing to verify is property (iv), which we will do assuming that is sufficiently small depending on
. From (3) we see that on
, we have (in the classical sense) that
and so (5) holds with forcing term
To show that is weakly infinitesimal, it suffices as before to show that
is strongly infinitesimal in for every standard
and compact
. But from (7) we know that
for all
,
, and
, and hence by overspill one has
for the same range of if
is sufficiently small depending on
. Thus
is strongly equivalent in
(and hence in
to the commutator type expression
But from Bernstein’s inequality and the boundedness of
, we know that
is strongly equivalent to
in
, so by Hölder it suffices to show that
is strongly infinitesimal in
. But this follows from the fact that
is strongly equivalent to
in
, and that
vanishes.
Remark 17 The above proof shows that we can in fact demand stronger regularity on the time derivative
than is required in Proposition 16(iii) if desired; for instance, one can place
in
for
close enough to
.
Exercise 18 State and prove a more traditional analogue of this proposition that asserts (roughly speaking) that any weak limit of a sequence of smooth solutions to Navier-Stokes with changes in initial data and forcing term that converge weakly to zero, which asymptotically obeys the energy inequality, and which has some weak uniform control on the time derivative, will produce a Leray-Hopf solution, and conversely that every Leray-Hopf solution arises in this fashion.
Exercise 19 Translate the proof of weak-strong uniqueness from Proposition 20 of Notes 2 to nonstandard analysis, by first using Proposition 16 to interpret the weak solution as the weakly standard part of a strong nonstandard approximate solution. (One will need the improved control on
mentioned in Remark 6.)
6 comments
Comments feed for this article
9 December, 2018 at 6:55 pm
Anonymous
Is there a simple characterization of the cases in which it is simpler to work with nonstandard analysis ?
10 December, 2018 at 7:57 am
Terence Tao
Well, as I see it nonstandard analysis offers two main advantages: (a) it cleans up a lot of the “epsilon management“, in that one has to keep less track of whether (say) epsilon depends on delta or vice versa, and (b) it allows one to directly apply continuous or infinitary mathematics (e.g., ergodic theory, topological group theory, algebraic geometry) to discrete or finitary mathematics (e.g., combinatorics). (As mentioned here, one also has (c) less reliance on any sort of “countability” or “separability” hypothesis on the spaces being considered, though this is a more minor advantage, for instance it can be largely duplicated in sequential analysis by switching from sequences to nets.)
(or even worse, growth function type parameters
, as is the case for instance with modern regularity lemmas), and (b) is “trying” to use some poorly developed discrete analogue of an existing continuous mathematical theory (e.g., “approximate ergodic theory”, “approximate topological group theory”, etc.). For instance my work with Breuillard and Green (building on earlier work of Hrushovski) on classifying approximate subgroups in discrete groups by using the theory of Hilbert’s fifth problem in topological group theory is a good example of such a problem (indeed currently there is no proof of this classification that does not go through nonstandard analysis or something closely related to it).
So a good candidate for a problem which could be simplified by a nonstandard approach would be a discrete problem in which the traditional attack on the problem (a) contains plenty of epsilon type parameters
In the case of PDE, several of the epsilons and deltas can already be concealed using the machinery of sequences converging in various strong or weak topologies (though, as I mentioned in a different post, this machinery struggles a bit more with more advanced notions such as concentration compactness), and of course PDE are already continuous instead of discrete. So admittedly the advantages of a nonstandard formalism here are rather slight (basically one is only left with (c), and also one no longer needs to keep passing to subsequences all the time). But I found it an interesting exercise to perform the translation anyway, as it led to some ways of thinking about real numbers, distributions and weak solutions that I had not fully appreciated before. Namely, in the nonstandard formalism, real numbers are the standard part of bounded nonstandard rationals, distributions are the weakly standard part of weakly bounded nonstandard test functions, and weak solutions are similarly the weakly standard part of strong approximate nonstandard solutions (with weakly infinitesimal errors) that obey some low-regularity bounds. Back in the usual sequential approach to mathematics, I knew of the analogous principles for the first two assertions (namely, reals are the limits of sequences of rationals, and distributions are the weak limits of sequences of test functions), but was a bit hazy on the third (namely, that weak solutions are always the weak limit of sequences of strong approximate solutions with errors that go weakly to zero).
Finally, as mentioned in the post, this exercise gave me a glimpse into a possible “alternate history” of the development of mathematics, in which analysis was largely formulated around nonstandard concepts such as infinitesimals instead of the orthodox approach using epsilons, deltas, and limits of sequences. In this particular case it seems this alternate foundation is largely equivalent to the one we actually have, though.
13 December, 2018 at 9:53 pm
Andrew Warren
Dear Terry,
You may already have run across this, but there is a rather nice monograph in this vein by Capinski and Cutland called “Nonstandard methods in stochastic fluid mechanics”. As the title suggests, they make good use of the Loeb space formulation of Browian motion/Ito calculus/etc.
16 December, 2018 at 6:49 am
Anonymous
Dear Terry,
If Sarnak and Chowla conjecture is proved completely 100% on time this Christmas holiday.How great it is! I am expecting from your big Noel gift(a gift for math
community not only for me)
18 December, 2018 at 4:33 am
GWCurvatureFlow
Dear Terry,
Just a quick question — do you know of any reference (perhaps have done it already yourself somewhere?) to prove well-posedness for nonlinear elliptic and/or parabolic PDE with nonstandard analysis?
Would be interesting to have a nonstandard treatment!
Best,
Glen
4 January, 2021 at 6:59 am
Go 2 It
1. I wasn’t able to verify the claim, stated following Exercise 9, that for a standard Hilbert space
with basis
and an unbounded natural number
, the vector
lies in
. Let
be defined by
, where
vanishes
is a power of two, in which case
, so that
is finite. Take for
an unbounded power of two. Then
is unbounded. More generally, it seems that Banach–Steinhaus gives
for a standard Banach space
, and in particular
for a standard Hilbert space
. Did I miss something?
Typo in the same paragraph: “weak*-bounded elements of
” should presumably be “weak*-bounded elements of
.”
[Typo and example corrected. Interestingly, the analogue of the Banach-Steinhaus theorem in this nonstandard setting appears to be false! -T]
2. In Exercise 13, it seems that
is missing, as we could otherwise take
times the characteristic function of a standard compact set of unit volume, with
a nonzero infinitesimal.
[Here we are using the seminorm that is the standard part of the nonstandard norm – see the first display of Section 2. -T]