Spaces of test functions and distributions

Summary

In mathematical analysis, the spaces of test functions and distributions are topological vector spaces (TVSs) that are used in the definition and application of distributions. Test functions are usually infinitely differentiable complex-valued (or sometimes real-valued) functions on a non-empty open subset that have compact support. The space of all test functions, denoted by is endowed with a certain topology, called the canonical LF-topology, that makes into a complete Hausdorff locally convex TVS. The strong dual space of is called the space of distributions on and is denoted by where the "" subscript indicates that the continuous dual space of denoted by is endowed with the strong dual topology.

There are other possible choices for the space of test functions, which lead to other different spaces of distributions. If then the use of Schwartz functions[note 1] as test functions gives rise to a certain subspace of whose elements are called tempered distributions. These are important because they allow the Fourier transform to be extended from "standard functions" to tempered distributions. The set of tempered distributions forms a vector subspace of the space of distributions and is thus one example of a space of distributions; there are many other spaces of distributions.

There also exist other major classes of test functions that are not subsets of such as spaces of analytic test functions, which produce very different classes of distributions. The theory of such distributions has a different character from the previous one because there are no analytic functions with non-empty compact support.[note 2] Use of analytic test functions leads to Sato's theory of hyperfunctions.

Notation edit

The following notation will be used throughout this article:

  •   is a fixed positive integer and   is a fixed non-empty open subset of Euclidean space  
  •   denotes the natural numbers.
  •   will denote a non-negative integer or  
  • If   is a function then   will denote its domain and the support of   denoted by   is defined to be the closure of the set   in  
  • For two functions  , the following notation defines a canonical pairing:
     
  • A multi-index of size   is an element in   (given that   is fixed, if the size of multi-indices is omitted then the size should be assumed to be  ). The length of a multi-index   is defined as   and denoted by   Multi-indices are particularly useful when dealing with functions of several variables, in particular we introduce the following notations for a given multi-index  :
     
    We also introduce a partial order of all multi-indices by   if and only if   for all   When   we define their multi-index binomial coefficient as:
     
  •   will denote a certain non-empty collection of compact subsets of   (described in detail below).

Definitions of test functions and distributions edit

In this section, we will formally define real-valued distributions on U. With minor modifications, one can also define complex-valued distributions, and one can replace   with any (paracompact) smooth manifold.

Notation:
  1. Let  
  2. Let   denote the vector space of all k-times continuously differentiable real or complex-valued functions on U.
  3. For any compact subset   let   and   both denote the vector space of all those functions   such that  
    • If   then the domain of   is U and not K. So although   depends on both K and U, only K is typically indicated. The justification for this common practice is detailed below. The notation   will only be used when the notation   risks being ambiguous.
    • Every   contains the constant 0 map, even if  
  4. Let   denote the set of all   such that   for some compact subset K of U.
    • Equivalently,   is the set of all   such that   has compact support.
    •   is equal to the union of all   as   ranges over  
    • If   is a real-valued function on U, then   is an element of   if and only if   is a   bump function. Every real-valued test function on   is always also a complex-valued test function on  
 
The graph of the bump function   where   and   This function is a test function on   and is an element of   The support of this function is the closed unit disk in   It is non-zero on the open unit disk and it is equal to 0 everywhere outside of it.

Note that for all   and any compact subsets K and L of U, we have:

 
Definition: Elements of   are called test functions on U and   is called the space of test function on U. We will use both   and   to denote this space.

Distributions on U are defined to be the continuous linear functionals on   when this vector space is endowed with a particular topology called the canonical LF-topology. This topology is unfortunately not easy to define but it is nevertheless still possible to characterize distributions in a way so that no mention of the canonical LF-topology is made.

Proposition: If T is a linear functional on   then the T is a distribution if and only if the following equivalent conditions are satisfied:

  1. For every compact subset   there exist constants   and   (dependent on  ) such that for all  [1]
     
  2. For every compact subset   there exist constants   and   such that for all   with support contained in  [2]
     
  3. For any compact subset   and any sequence   in   if   converges uniformly to zero on   for all multi-indices  , then  

The above characterizations can be used to determine whether or not a linear functional is a distribution, but more advanced uses of distributions and test functions (such as applications to differential equations) is limited if no topologies are placed on   and   To define the space of distributions we must first define the canonical LF-topology, which in turn requires that several other locally convex topological vector spaces (TVSs) be defined first. First, a (non-normable) topology on   will be defined, then every   will be endowed with the subspace topology induced on it by   and finally the (non-metrizable) canonical LF-topology on   will be defined. The space of distributions, being defined as the continuous dual space of   is then endowed with the (non-metrizable) strong dual topology induced by   and the canonical LF-topology (this topology is a generalization of the usual operator norm induced topology that is placed on the continuous dual spaces of normed spaces). This finally permits consideration of more advanced notions such as convergence of distributions (both sequences and nets), various (sub)spaces of distributions, and operations on distributions, including extending differential equations to distributions.

Choice of compact sets K edit

Throughout,   will be any collection of compact subsets of   such that (1)   and (2) for any compact   there exists some   such that   The most common choices for   are:

  • The set of all compact subsets of   or
  • A set   where   and for all i,   and   is a relatively compact non-empty open subset of   (here, "relatively compact" means that the closure of   in either U or   is compact).

We make   into a directed set by defining   if and only if   Note that although the definitions of the subsequently defined topologies explicitly reference   in reality they do not depend on the choice of   that is, if   and   are any two such collections of compact subsets of   then the topologies defined on   and   by using   in place of   are the same as those defined by using   in place of  

Topology on Ck(U) edit

We now introduce the seminorms that will define the topology on   Different authors sometimes use different families of seminorms so we list the most common families below. However, the resulting topology is the same no matter which family is used.

Suppose   and   is an arbitrary compact subset of   Suppose   an integer such that  [note 3] and   is a multi-index with length   For   define:

 
while for   define all the functions above to be the constant 0 map.

All of the functions above are non-negative  -valued[note 4] seminorms on   As explained in this article, every set of seminorms on a vector space induces a locally convex vector topology.

Each of the following sets of seminorms

 
generate the same locally convex vector topology on   (so for example, the topology generated by the seminorms in   is equal to the topology generated by those in  ).
The vector space   is endowed with the locally convex topology induced by any one of the four families   of seminorms described above. This topology is also equal to the vector topology induced by all of the seminorms in  

With this topology,   becomes a locally convex Fréchet space that is not normable. Every element of   is a continuous seminorm on   Under this topology, a net   in   converges to   if and only if for every multi-index   with   and every compact   the net of partial derivatives   converges uniformly to   on  [3] For any   any (von Neumann) bounded subset of   is a relatively compact subset of  [4] In particular, a subset of   is bounded if and only if it is bounded in   for all  [4] The space   is a Montel space if and only if  [5]

The topology on   is the superior limit of the subspace topologies induced on   by the TVSs   as i ranges over the non-negative integers.[3] A subset   of   is open in this topology if and only if there exists   such that   is open when   is endowed with the subspace topology induced on it by  

Metric defining the topology edit

If the family of compact sets   satisfies   and   for all   then a complete translation-invariant metric on   can be obtained by taking a suitable countable Fréchet combination of any one of the above defining families of seminorms (A through D). For example, using the seminorms   results in the metric

 

Often, it is easier to just consider seminorms (avoiding any metric) and use the tools of functional analysis.

Topology on Ck(K) edit

As before, fix   Recall that if   is any compact subset of   then  

Assumption: For any compact subset   we will henceforth assume that   is endowed with the subspace topology it inherits from the Fréchet space  

For any compact subset     is a closed subspace of the Fréchet space   and is thus also a Fréchet space. For all compact   satisfying   denote the inclusion map by   Then this map is a linear embedding of TVSs (that is, it is a linear map that is also a topological embedding) whose image (or "range") is closed in its codomain; said differently, the topology on   is identical to the subspace topology it inherits from   and also   is a closed subset of   The interior of   relative to   is empty.[6]

If   is finite then   is a Banach space[7] with a topology that can be defined by the norm

 

And when   then   is even a Hilbert space.[7] The space   is a distinguished Schwartz Montel space so if   then it is not normable and thus not a Banach space (although like all other   it is a Fréchet space).

Trivial extensions and independence of Ck(K)'s topology from U edit

The definition of   depends on U so we will let   denote the topological space   which by definition is a topological subspace of   Suppose   is an open subset of   containing   and for any compact subset   let   is the vector subspace of   consisting of maps with support contained in   Given   its trivial extension to V is by definition, the function   defined by:

 
so that   Let   denote the map that sends a function in   to its trivial extension on V. This map is a linear injection and for every compact subset   (where   is also a compact subset of   since  ) we have
 
If I is restricted to   then the following induced linear map is a homeomorphism (and thus a TVS-isomorphism):
 
and thus the next two maps (which like the previous map are defined by  ) are topological embeddings:
 
(the topology on   is the canonical LF topology, which is defined later). Using the injection
 
the vector space   is canonically identified with its image in   (however, if   then   is not a topological embedding when these spaces are endowed with their canonical LF topologies, although it is continuous).[8] Because   through this identification,   can also be considered as a subset of   Importantly, the subspace topology   inherits from   (when it is viewed as a subset of  ) is identical to the subspace topology that it inherits from   (when   is viewed instead as a subset of   via the identification). Thus the topology on   is independent of the open subset U of   that contains K.[6] This justifies the practice of written   instead of  

Canonical LF topology edit

Recall that   denote all those functions in   that have compact support in   where note that   is the union of all   as K ranges over   Moreover, for every k,   is a dense subset of   The special case when   gives us the space of test functions.

  is called the space of test functions on   and it may also be denoted by  

This section defines the canonical LF topology as a direct limit. It is also possible to define this topology in terms of its neighborhoods of the origin, which is described afterwards.

Topology defined by direct limits edit

For any two sets K and L, we declare that   if and only if   which in particular makes the collection   of compact subsets of U into a directed set (we say that such a collection is directed by subset inclusion). For all compact   satisfying   there are inclusion maps

 

Recall from above that the map   is a topological embedding. The collection of maps

 
forms a direct system in the category of locally convex topological vector spaces that is directed by   (under subset inclusion). This system's direct limit (in the category of locally convex TVSs) is the pair   where   are the natural inclusions and where   is now endowed with the (unique) strongest locally convex topology making all of the inclusion maps   continuous.
The canonical LF topology on   is the finest locally convex topology on   making all of the inclusion maps   continuous (where K ranges over  ).
As is common in mathematics literature, the space   is henceforth assumed to be endowed with its canonical LF topology (unless explicitly stated otherwise).

Topology defined by neighborhoods of the origin edit

If U is a convex subset of   then U is a neighborhood of the origin in the canonical LF topology if and only if it satisfies the following condition:

For all     is a neighborhood of the origin in  

(CN)

Note that any convex set satisfying this condition is necessarily absorbing in   Since the topology of any topological vector space is translation-invariant, any TVS-topology is completely determined by the set of neighborhood of the origin. This means that one could actually define the canonical LF topology by declaring that a convex balanced subset U is a neighborhood of the origin if and only if it satisfies condition CN.

Topology defined via differential operators edit

A linear differential operator in U with smooth coefficients is a sum

 
where   and all but finitely many of   are identically 0. The integer   is called the order of the differential operator   If   is a linear differential operator of order k then it induces a canonical linear map   defined by   where we shall reuse notation and also denote this map by  [9]

For any   the canonical LF topology on   is the weakest locally convex TVS topology making all linear differential operators in   of order   into continuous maps from   into  [9]

Properties of the canonical LF topology edit

Canonical LF topology's independence from K edit

One benefit of defining the canonical LF topology as the direct limit of a direct system is that we may immediately use the universal property of direct limits. Another benefit is that we can use well-known results from category theory to deduce that the canonical LF topology is actually independent of the particular choice of the directed collection   of compact sets. And by considering different collections   (in particular, those   mentioned at the beginning of this article), we may deduce different properties of this topology. In particular, we may deduce that the canonical LF topology makes   into a Hausdorff locally convex strict LF-space (and also a strict LB-space if  ), which of course is the reason why this topology is called "the canonical LF topology" (see this footnote for more details).[note 5]

Universal property edit

From the universal property of direct limits, we know that if   is a linear map into a locally convex space Y (not necessarily Hausdorff), then u is continuous if and only if u is bounded if and only if for every   the restriction of u to   is continuous (or bounded).[10][11]

Dependence of the canonical LF topology on U edit

Suppose V is an open subset of   containing   Let   denote the map that sends a function in   to its trivial extension on V (which was defined above). This map is a continuous linear map.[8] If (and only if)   then   is not a dense subset of   and   is not a topological embedding.[8] Consequently, if   then the transpose of   is neither one-to-one nor onto.[8]

Bounded subsets edit

A subset   is bounded in   if and only if there exists some   such that   and   is a bounded subset of  [11] Moreover, if   is compact and   then   is bounded in   if and only if it is bounded in   For any   any bounded subset of   (resp.  ) is a relatively compact subset of   (resp.  ), where  [11]

Non-metrizability edit

For all compact   the interior of   in   is empty so that   is of the first category in itself. It follows from Baire's theorem that   is not metrizable and thus also not normable (see this footnote[note 6] for an explanation of how the non-metrizable space   can be complete even though it does not admit a metric). The fact that   is a nuclear Montel space makes up for the non-metrizability of   (see this footnote for a more detailed explanation).[note 7]

Relationships between spaces edit

Using the universal property of direct limits and the fact that the natural inclusions   are all topological embedding, one may show that all of the maps   are also topological embeddings. Said differently, the topology on   is identical to the subspace topology that it inherits from   where recall that  's topology was defined to be the subspace topology induced on it by   In particular, both   and   induces the same subspace topology on   However, this does not imply that the canonical LF topology on   is equal to the subspace topology induced on   by  ; these two topologies on   are in fact never equal to each other since the canonical LF topology is never metrizable while the subspace topology induced on it by   is metrizable (since recall that   is metrizable). The canonical LF topology on   is actually strictly finer than the subspace topology that it inherits from   (thus the natural inclusion   is continuous but not a topological embedding).[7]

Indeed, the canonical LF topology is so fine that if   denotes some linear map that is a "natural inclusion" (such as   or   or other maps discussed below) then this map will typically be continuous, which (as is explained below) is ultimately the reason why locally integrable functions, Radon measures, etc. all induce distributions (via the transpose of such a "natural inclusion"). Said differently, the reason why there are so many different ways of defining distributions from other spaces ultimately stems from how very fine the canonical LF topology is. Moreover, since distributions are just continuous linear functionals on   the fine nature of the canonical LF topology means that more linear functionals on   end up being continuous ("more" means as compared to a coarser topology that we could have placed on   such as for instance, the subspace topology induced by some   which although it would have made   metrizable, it would have also resulted in fewer linear functionals on   being continuous and thus there would have been fewer distributions; moreover, this particular coarser topology also has the disadvantage of not making   into a complete TVS[12]).

Other properties edit
  • The differentiation map   is a surjective continuous linear operator.[13]
  • The bilinear multiplication map   given by   is not continuous; it is however, hypocontinuous.[14]

Distributions edit

As discussed earlier, continuous linear functionals on a   are known as distributions on U. Thus the set of all distributions on U is the continuous dual space of   which when endowed with the strong dual topology is denoted by  

By definition, a distribution on U is defined to be a continuous linear functional on   Said differently, a distribution on U is an element of the continuous dual space of   when   is endowed with its canonical LF topology.

We have the canonical duality pairing between a distribution T on U and a test function   which is denoted using angle brackets by

 

One interprets this notation as the distribution T acting on the test function   to give a scalar, or symmetrically as the test function   acting on the distribution T.

Characterizations of distributions edit

Proposition. If T is a linear functional on   then the following are equivalent:

  1. T is a distribution;
  2. Definition : T is a continuous function.
  3. T is continuous at the origin.
  4. T is uniformly continuous.
  5. T is a bounded operator.
  6. T is sequentially continuous.
    • explicitly, for every sequence   in   that converges in   to some    [note 8]
  7. T is sequentially continuous at the origin; in other words, T maps null sequences[note 9] to null sequences.
    • explicitly, for every sequence   in   that converges in   to the origin (such a sequence is called a null sequence),  
    • a null sequence is by definition a sequence that converges to the origin.
  8. T maps null sequences to bounded subsets.
    • explicitly, for every sequence   in   that converges in   to the origin, the sequence   is bounded.
  9. T maps Mackey convergent null sequences[note 10] to bounded subsets;
    • explicitly, for every Mackey convergent null sequence   in   the sequence   is bounded.
    • a sequence   is said to be Mackey convergent to 0 if there exists a divergent sequence   of positive real number such that the sequence   is bounded; every sequence that is Mackey convergent to 0 necessarily converges to the origin (in the usual sense).
  10. The kernel of T is a closed subspace of  
  11. The graph of T is closed.
  12. There exists a continuous seminorm   on   such that  
  13. There exists a constant   a collection of continuous seminorms,   that defines the canonical LF topology of   and a finite subset   such that  [note 11]
  14. For every compact subset   there exist constants   and   such that for all  [1]
     
  15. For every compact subset   there exist constants   and   such that for all   with support contained in  [2]
     
  16. For any compact subset   and any sequence   in   if   converges uniformly to zero for all multi-indices   then  
  17. Any of the three statements immediately above (that is, statements 14, 15, and 16) but with the additional requirement that compact set   belongs to  

Topology on the space of distributions edit

Definition and notation: The space of distributions on U, denoted by   is the continuous dual space of   endowed with the topology of uniform convergence on bounded subsets of  [7] More succinctly, the space of distributions on U is  

The topology of uniform convergence on bounded subsets is also called the strong dual topology.[note 12] This topology is chosen because it is with this topology that   becomes a nuclear Montel space and it is with this topology that the kernels theorem of Schwartz holds.[15] No matter what dual topology is placed on  [note 13] a sequence of distributions converges in this topology if and only if it converges pointwise (although this need not be true of a net). No matter which topology is chosen,   will be a non-metrizable, locally convex topological vector space. The space   is separable[16] and has the strong Pytkeev property[17] but it is neither a k-space[17] nor a sequential space,[16] which in particular implies that it is not metrizable and also that its topology can not be defined using only sequences.

Topological properties edit

Topological vector space categories edit

The canonical LF topology makes   into a complete distinguished strict LF-space (and a strict LB-space if and only if  [18]), which implies that   is a meager subset of itself.[19] Furthermore,   as well as its strong dual space, is a complete Hausdorff locally convex barrelled bornological Mackey space. The strong dual of   is a Fréchet space if and only if   so in particular, the strong dual of   which is the space   of distributions on U, is not metrizable (note that the weak-* topology on   also is not metrizable and moreover, it further lacks almost all of the nice properties that the strong dual topology gives  ).

The three spaces     and the Schwartz space   as well as the strong duals of each of these three spaces, are complete nuclear[20] Montel[21] bornological spaces, which implies that all six of these locally convex spaces are also paracompact[22] reflexive barrelled Mackey spaces. The spaces   and   are both distinguished Fréchet spaces. Moreover, both   and   are Schwartz TVSs.

Convergent sequences edit

Convergent sequences and their insufficiency to describe topologies edit

The strong dual spaces of   and   are sequential spaces but not Fréchet-Urysohn spaces.[16] Moreover, neither the space of test functions   nor its strong dual   is a sequential space (not even an Ascoli space),[16][23] which in particular implies that their topologies can not be defined entirely in terms of convergent sequences.

A sequence   in   converges in   if and only if there exists some   such that   contains this sequence and this sequence converges in  ; equivalently, it converges if and only if the following two conditions hold:[24]

  1. There is a compact set   containing the supports of all  
  2. For each multi-index   the sequence of partial derivatives   tends uniformly to  

Neither the space   nor its strong dual   is a sequential space,[16][23] and consequently, their topologies can not be defined entirely in terms of convergent sequences. For this reason, the above characterization of when a sequence converges is not enough to define the canonical LF topology on   The same can be said of the strong dual topology on  

What sequences do characterize edit

Nevertheless, sequences do characterize many important properties, as we now discuss. It is known that in the dual space of any Montel space, a sequence converges in the strong dual topology if and only if it converges in the weak* topology,[25] which in particular, is the reason why a sequence of distributions converges (in the strong dual topology) if and only if it converges pointwise (this leads many authors to use pointwise convergence to actually define the convergence of a sequence of distributions; this is fine for sequences but it does not extend to the convergence of nets of distributions since a net may converge pointwise but fail to converge in the strong dual topology).

Sequences characterize continuity of linear maps valued in locally convex space. Suppose X is a locally convex bornological space (such as any of the six TVSs mentioned earlier). Then a linear map   into a locally convex space Y is continuous if and only if it maps null sequences[note 9] in X to bounded subsets of Y.[note 14] More generally, such a linear map   is continuous if and only if it maps Mackey convergent null sequences[note 10] to bounded subsets of   So in particular, if a linear map   into a locally convex space is sequentially continuous at the origin then it is continuous.[26] However, this does not necessarily extend to non-linear maps and/or to maps valued in topological spaces that are not locally convex TVSs.

For every   is sequentially dense in  [27] Furthermore,   is a sequentially dense subset of   (with its strong dual topology)[28] and also a sequentially dense subset of the strong dual space of  [28]

Sequences of distributions edit

A sequence of distributions   converges with respect to the weak-* topology on   to a distribution T if and only if

 
for every test function   For example, if   is the function
 
and   is the distribution corresponding to   then
 
as   so   in   Thus, for large   the function   can be regarded as an approximation of the Dirac delta distribution.
Other properties edit
  • The strong dual space of   is TVS isomorphic to   via the canonical TVS-isomorphism   defined by sending   to value at   (that is, to the linear functional on   defined by sending   to  );
  • On any bounded subset of   the weak and strong subspace topologies coincide; the same is true for  ;
  • Every weakly convergent sequence in   is strongly convergent (although this does not extend to nets).

Localization of distributions edit

Preliminaries: Transpose of a linear operator edit

Operations on distributions and spaces of distributions are often defined by means of the transpose of a linear operator. This is because the transpose allows for a unified presentation of the many definitions in the theory of distributions and also because its properties are well known in functional analysis.[29] For instance, the well-known Hermitian adjoint of a linear operator between Hilbert spaces is just the operator's transpose (but with the Riesz representation theorem used to identify each Hilbert space with its continuous dual space). In general the transpose of a continuous linear map   is the linear map

 
or equivalently, it is the unique map satisfying   for all   and all   (the prime symbol in   does not denote a derivative of any kind; it merely indicates that   is an element of the continuous dual space  ). Since   is continuous, the transpose   is also continuous when both duals are endowed with their respective strong dual topologies; it is also continuous when both duals are endowed with their respective weak* topologies (see the articles polar topology and dual system for more details).

In the context of distributions, the characterization of the transpose can be refined slightly. Let   be a continuous linear map. Then by definition, the transpose of   is the unique linear operator   that satisfies:

 

Since   is dense in   (here,   actually refers to the set of distributions  ) it is sufficient that the defining equality hold for all distributions of the form   where   Explicitly, this means that a continuous linear map   is equal to   if and only if the condition below holds:

 
where the right hand side equals  

Extensions and restrictions to an open subset edit

Let   be open subsets of   Every function   can be extended by zero from its domain   to a function on   by setting it equal to   on the complement   This extension is a smooth compactly supported function called the trivial extension of   to   and it will be denoted by   This assignment   defines the trivial extension operator   which is a continuous injective linear map. It is used to canonically identify   as a vector subspace of   (although not as a topological subspace). Its transpose (explained here)

 
is called the restriction to   of distributions in  [8] and as the name suggests, the image   of a distribution   under this map is a distribution on   called the restriction of   to   The defining condition of the restriction   is:
 
If   then the (continuous injective linear) trivial extension map   is not a topological embedding (in other words, if this linear injection was used to identify   as a subset of   then  's topology would strictly finer than the subspace topology that   induces on it; importantly, it would not be a topological subspace since that requires equality of topologies) and its range is also not dense in its codomain  [8] Consequently, if   then the restriction mapping is neither injective nor surjective.[8] A distribution   is said to be extendible to U if it belongs to the range of the transpose of   and it is called extendible if it is extendable to  [8]

Unless   the restriction to   is neither injective nor surjective.

Spaces of distributions edit

For all   and all   all of the following canonical injections are continuous and have an image/range that is a dense subset of their codomain:[30][31]

 
where the topologies on the LB-spaces   are the canonical LF topologies as defined below (so in particular, they are not the usual norm topologies). The range of each of the maps above (and of any composition of the maps above) is dense in the codomain. Indeed,   is even sequentially dense in every  [27] For every   the canonical inclusion   into the normed space   (here   has its usual norm topology) is a continuous linear injection and the range of this injection is dense in its codomain if and only if   .[31]

Suppose that   is one of the LF-spaces   (for  ) or LB-spaces   (for  ) or normed spaces   (for  ).[31] Because the canonical injection   is a continuous injection whose image is dense in the codomain, this map's transpose   is a continuous injection. This injective transpose map thus allows the continuous dual space   of   to be identified with a certain vector subspace of the space   of all distributions (specifically, it is identified with the image of this transpose map). This continuous transpose map is not necessarily a TVS-embedding so the topology that this map transfers from its domain to the image   is finer than the subspace topology that this space inherits from   A linear subspace of   carrying a locally convex topology that is finer than the subspace topology induced by   is called a space of distributions.[32] Almost all of the spaces of distributions mentioned in this article arise in this way (e.g. tempered distribution, restrictions, distributions of order   some integer, distributions induced by a positive Radon measure, distributions induced by an  -function, etc.) and any representation theorem about the dual space of X may, through the transpose   be transferred directly to elements of the space  

Compactly supported Lp-spaces edit

Given   the vector space   of compactly supported   functions on   and its topology are defined as direct limits of the spaces   in a manner analogous to how the canonical LF-topologies on   were defined. For any compact   let   denote the set of all element in   (which recall are equivalence class of Lebesgue measurable   functions on  ) having a representative   whose support (which recall is the closure of   in  ) is a subset of   (such an   is almost everywhere defined in  ). The set   is a closed vector subspace   and is thus a Banach space and when   even a Hilbert space.[30] Let   be the union of all   as   ranges over all compact subsets of   The set   is a vector subspace of   whose elements are the (equivalence classes of) compactly supported   functions defined on   (or almost everywhere on  ). Endow   with the final topology (direct limit topology) induced by the inclusion maps   as   ranges over all compact subsets of   This topology is called the canonical LF topology and it is equal to the final topology induced by any countable set of inclusion maps   ( ) where   are any compact sets with union equal to  [30] This topology makes   into an LB-space (and thus also an LF-space) with a topology that is strictly finer than the norm (subspace) topology that   induces on it.

Radon measures edit

The inclusion map   is a continuous injection whose image is dense in its codomain, so the transpose   is also a continuous injection.

Note that the continuous dual space   can be identified as the space of Radon measures, where there is a one-to-one correspondence between the continuous linear functionals   and integral with respect to a Radon measure; that is,

  • if   then there exists a Radon measure   on U such that for all   and
  • if   is a Radon measure on U then the linear functional on   defined by   is continuous.

Through the injection   every Radon measure becomes a distribution on U. If   is a locally integrable function on U then the distribution   is a Radon measure; so Radon measures form a large and important space of distributions.

The following is the theorem of the structure of distributions of Radon measures, which shows that every Radon measure can be written as a sum of derivatives of locally   functions in U :

Theorem.[33] — Suppose   is a Radon measure, where   let   be a neighborhood of the support of   and let   There exists a family   of locally   functions on U such that   for every   and

 
Furthermore,   is also equal to a finite sum of derivatives of continuous functions on   where each derivative has order  

Positive Radon measures

A linear function T on a space of functions is called positive if whenever a function   that belongs to the domain of T is non-negative (meaning that   is real-valued and  ) then   One may show that every positive linear functional on   is necessarily continuous (that is, necessarily a Radon measure).[34] Lebesgue measure is an example of a positive Radon measure.

Locally integrable functions as distributions edit

One particularly important class of Radon measures are those that are induced locally integrable functions. The function   is called locally integrable if it is Lebesgue integrable over every compact subset K of U.[note 15] This is a large class of functions which includes all continuous functions and all Lp space   functions. The topology on   is defined in such a fashion that any locally integrable function   yields a continuous linear functional on   – that is, an element of   – denoted here by  , whose value on the test function   is given by the Lebesgue integral:

 

Conventionally, one abuses notation by identifying   with   provided no confusion can arise, and thus the pairing between   and   is often written

 

If   and g are two locally integrable functions, then the associated distributions   and Tg are equal to the same element of   if and only if   and g are equal almost everywhere (see, for instance, Hörmander (1983, Theorem 1.2.5)). In a similar manner, every Radon measure   on U defines an element of   whose value on the test function   is   As above, it is conventional to abuse notation and write the pairing between a Radon measure   and a test function   as   Conversely, as shown in a theorem by Schwartz (similar to the Riesz representation theorem), every distribution which is non-negative on non-negative functions is of this form for some (positive) Radon measure.

Test functions as distributions

The test functions are themselves locally integrable, and so define distributions. The space of test functions   is sequentially dense in   with respect to the strong topology on  [28] This means that for any   there is a sequence of test functions,   that converges to   (in its strong dual topology) when considered as a sequence of distributions. Or equivalently,

 

Furthermore,   is also sequentially dense in the strong dual space of  [28]

Distributions with compact support edit

The inclusion map   is a continuous injection whose image is dense in its codomain, so the transpose   is also a continuous injection. Thus the image of the transpose, denoted by   forms a space of distributions when it is endowed with the strong dual topology of   (transferred to it via the transpose map   so the topology of   is finer than the subspace topology that this set inherits from  ).[35]

The elements of   can be identified as the space of distributions with compact support.[35] Explicitly, if T is a distribution on U then the following are equivalent,

  •  ;
  • the support of T is compact;
  • the restriction of   to   when that space is equipped with the subspace topology inherited from   (a coarser topology than the canonical LF topology), is continuous;[35]
  • there is a compact subset K of U such that for every test function   whose support is completely outside of K, we have  

Compactly supported distributions define continuous linear functionals on the space  ; recall that the topology on   is defined such that a sequence of test functions   converges to 0 if and only if all derivatives of   converge uniformly to 0 on every compact subset of U. Conversely, it can be shown that every continuous linear functional on this space defines a distribution of compact support. Thus compactly supported distributions can be identified with those distributions that can be extended from   to  

Distributions of finite order edit

Let   The inclusion map   is a continuous injection whose image is dense in its codomain, so the transpose   is also a continuous injection. Consequently, the image of   denoted by   forms a space of distributions when it is endowed with the strong dual topology of   (transferred to it via the transpose map   so  's topology is finer than the subspace topology that this set inherits from  ). The elements of   are the distributions of order  [36] The distributions of order   which are also called distributions of order   are exactly the distributions that are Radon measures (described above).

For   a distribution of order   is a distribution of order   that is not a distribution of order  [36]

A distribution is said to be of finite order if there is some integer k such that it is a distribution of order   and the set of distributions of finite order is denoted by   Note that if   then   so that   is a vector subspace of   and furthermore, if and only if  [36]

Structure of distributions of finite order

Every distribution with compact support in U is a distribution of finite order.[36] Indeed, every distribution in U is locally a distribution of finite order, in the following sense:[36] If V is an open and relatively compact subset of U and if   is the restriction mapping from U to V, then the image of   under   is contained in  

The following is the theorem of the structure of distributions of finite order, which shows that every distribution of finite order can be written as a sum of derivatives of Radon measures:

Theorem[36] — Suppose   has finite order and   Given any open subset V of U containing the support of T, there is a family of Radon measures in U,   such that for very   and

 

Example. (Distributions of infinite order) Let   and for every test function   let

 

Then S is a distribution of infinite order on U. Moreover, S can not be extended to a distribution on  ; that is, there exists no distribution T on   such that the restriction of T to U is equal to T.[37]

Tempered distributions and Fourier transform edit

Defined below are the tempered distributions, which form a subspace of   the space of distributions on   This is a proper subspace: while every tempered distribution is a distribution and an element of   the converse is not true. Tempered distributions are useful if one studies the Fourier transform since all tempered distributions have a Fourier transform, which is not true for an arbitrary distribution in  

Schwartz space

The Schwartz space,   is the space of all smooth functions that are rapidly decreasing at infinity along with all partial derivatives. Thus   is in the Schwartz space provided that any derivative of   multiplied with any power of   converges to 0 as   These functions form a complete TVS with a suitably defined family of seminorms. More precisely, for any multi-indices   and   define:

 

Then   is in the Schwartz space if all the values satisfy:

 

The family of seminorms   defines a locally convex topology on the Schwartz space. For   the seminorms are, in fact, norms on the Schwartz space. One can also use the following family of seminorms to define the topology:[38]

 

Otherwise, one can define a norm on   via

 

The Schwartz space is a Fréchet space (i.e. a complete metrizable locally convex space). Because the Fourier transform changes   into multiplication by   and vice versa, this symmetry implies that the Fourier transform of a Schwartz function is also a Schwartz function.

A sequence   in   converges to 0 in   if and only if the functions   converge to 0 uniformly in the whole of   which implies that such a sequence must converge to zero in  [38]

  is dense in   The subset of all analytic Schwartz functions is dense in   as well.[39]

The Schwartz space is nuclear and the tensor product of two maps induces a canonical surjective TVS-isomorphisms

 
where   represents the completion of the injective tensor product (which in this case is the identical to the completion of the projective tensor product).[40]

Tempered distributions

The inclusion map   is a continuous injection whose image is dense in its codomain, so the transpose   is also a continuous injection. Thus, the image of the transpose map, denoted by   forms a space of distributions when it is endowed with the strong dual topology of   (transferred to it via the transpose map   so the topology of   is finer than the subspace topology that this set inherits from  ).

The space   is called the space of tempered distributions. It is the continuous dual of the Schwartz space. Equivalently, a distribution T is a tempered distribution if and only if

 

The derivative of a tempered distribution is again a tempered distribution. Tempered distributions generalize the bounded (or slow-growing) locally integrable functions; all distributions with compact support and all square-integrable functions are tempered distributions. More generally, all functions that are products of polynomials with elements of Lp space   for   are tempered distributions.

The tempered distributions can also be characterized as slowly growing, meaning that each derivative of T grows at most as fast as some polynomial. This characterization is dual to the rapidly falling behaviour of the derivatives of a function in the Schwartz space, where each derivative of   decays faster than every inverse power of   An example of a rapidly falling function is   for any positive  

Fourier transform

To study the Fourier transform, it is best to consider complex-valued test functions and complex-linear distributions. The ordinary continuous Fourier transform   is a TVS-automorphism of the Schwartz space, and the Fourier transform is defined to be its transpose   which (abusing notation) will again be denoted by F. So the Fourier transform of the tempered distribution T is defined by   for every Schwartz function     is thus again a tempered distribution. The Fourier transform is a TVS isomorphism from the space of tempered distributions onto itself. This operation is compatible with differentiation in the sense that

 
and also with convolution: if T is a tempered distribution and   is a slowly increasing smooth function on     is again a tempered distribution and
 
is the convolution of   and  . In particular, the Fourier transform of the constant function equal to 1 is the   distribution.

Expressing tempered distributions as sums of derivatives

If   is a tempered distribution, then there exists a constant   and positive integers M and N such that for all Schwartz functions  

 

This estimate along with some techniques from functional analysis can be used to show that there is a continuous slowly increasing function F and a multi-index   such that

 

Restriction of distributions to compact sets

If   then for any compact set   there exists a continuous function F compactly supported in   (possibly on a larger set than K itself) and a multi-index   such that   on  

Tensor product of distributions edit

Let   and   be open sets. Assume all vector spaces to be over the field   where   or   For   define for every   and every   the following functions:

 

Given   and   define the following functions:

 
where   and   These definitions associate every   and   with the (respective) continuous linear map:
 

Moreover, if either   (resp.  ) has compact support then it also induces a continuous linear map of   (resp.  ).[41]

Fubini's theorem for distributions[41] — Let   and   If   then

 

The tensor product of   and   denoted by   or   is the distribution in   defined by:[41]

 

Schwartz kernel theorem edit

The tensor product defines a bilinear map

 
the span of the range of this map is a dense subspace of its codomain. Furthermore,  [41] Moreover   induces continuous bilinear maps:
 
where   denotes the space of distributions with compact support and   is the Schwartz space of rapidly decreasing functions.[14]

Schwartz kernel theorem[40] — Each of the canonical maps below (defined in the natural way) are TVS isomorphisms:

 
Here   represents the completion of the injective tensor product (which in this case is identical to the completion of the projective tensor product, since these spaces are nuclear) and   has the topology of uniform convergence on bounded subsets.

This result does not hold for Hilbert spaces such as   and its dual space.[42] Why does such a result hold for the space of distributions and test functions but not for other "nice" spaces like the Hilbert space  ? This question led Alexander Grothendieck to discover nuclear spaces, nuclear maps, and the injective tensor product. He ultimately showed that it is precisely because   is a nuclear space that the Schwartz kernel theorem holds. Like Hilbert spaces, nuclear spaces may be thought as of generalizations of finite dimensional Euclidean space.

Using holomorphic functions as test functions edit

The success of the theory led to investigation of the idea of hyperfunction, in which spaces of holomorphic functions are used as test functions. A refined theory has been developed, in particular Mikio Sato's algebraic analysis, using sheaf theory and several complex variables. This extends the range of symbolic methods that can be made into rigorous mathematics, for example Feynman integrals.

See also edit

Notes edit

  1. ^ The Schwartz space consists of smooth rapidly decreasing test functions, where "rapidly decreasing" means that the function decreases faster than any polynomial increases as points in its domain move away from the origin.
  2. ^ Except for the trivial (i.e. identically  ) map, which of course is always analytic.
  3. ^ Note that   being an integer implies   This is sometimes expressed as   Since   the inequality " " means:   if   while if   then it means  
  4. ^ The image of the compact set   under a continuous  -valued map (for example,   for  ) is itself a compact, and thus bounded, subset of   If   then this implies that each of the functions defined above is  -valued (that is, none of the supremums above are ever equal to  ).
  5. ^ If we take   to be the set of all compact subsets of U then we can use the universal property of direct limits to conclude that the inclusion   is a continuous and even that they are topological embedding for every compact subset   If however, we take   to be the set of closures of some countable increasing sequence of relatively compact open subsets of U having all of the properties mentioned earlier in this in this article then we immediately deduce that   is a Hausdorff locally convex strict LF-space (and even a strict LB-space when  ). All of these facts can also be proved directly without using direct systems (although with more work).
  6. ^ For any TVS X (metrizable or otherwise), the notion of completeness depends entirely on a certain so-called "canonical uniformity" that is defined using only the subtraction operation (see the article Complete topological vector space for more details). In this way, the notion of a complete TVS does not require the existence of any metric. However, if the TVS X is metrizable and if   is any translation-invariant metric on X that defines its topology, then X is complete as a TVS (i.e. it is a complete uniform space under its canonical uniformity) if and only if   is a complete metric space. So if a TVS X happens to have a topology that can be defined by such a metric d then d may be used to deduce the completeness of X but the existence of such a metric is not necessary for defining completeness and it is even possible to deduce that a metrizable TVS is complete without ever even considering a metric (e.g. since the Cartesian product of any collection of complete TVSs is again a complete TVS, we can immediately deduce that the TVS   which happens to be metrizable, is a complete TVS; note that there was no need to consider any metric on  ).
  7. ^ One reason for giving   the canonical LF topology is because it is with this topology that   and its continuous dual space both become nuclear spaces, which have many nice properties and which may be viewed as a generalization of finite-dimensional spaces (for comparison, normed spaces are another generalization of finite-dimensional spaces that have many "nice" properties). In more detail, there are two classes of topological vector spaces (TVSs) that are particularly similar to finite-dimensional Euclidean spaces: the Banach spaces (especially Hilbert spaces) and the nuclear Montel spaces. Montel spaces are a class of TVSs in which every closed and bounded subset is compact (this generalizes the Heine–Borel theorem), which is a property that no infinite-dimensional Banach space can have; that is, no infinite-dimensional TVS can be both a Banach space and a Montel space. Also, no infinite-dimensional TVS can be both a Banach space and a nuclear space. All finite dimensional Euclidean spaces are nuclear Montel Hilbert spaces but once one enters infinite-dimensional space then these two classes separate. Nuclear spaces in particular have many of the "nice" properties of finite-dimensional TVSs (e.g. the Schwartz kernel theorem) that infinite-dimensional Banach spaces lack (for more details, see the properties, sufficient conditions, and characterizations given in the article Nuclear space). It is in this sense that nuclear spaces are an "alternative generalization" of finite-dimensional spaces. Also, as a general rule, in practice most "naturally occurring" TVSs are usually either Banach spaces or nuclear space. Typically, most TVSs that are associated with smoothness (i.e. infinite differentiability, such as   and  ) end up being nuclear TVSs while TVSs associated with finite continuous differentiability (such as   with K compact and  ) often end up being non-nuclear spaces, such as Banach spaces.
  8. ^ Even though the topology of   is not metrizable, a linear functional on   is continuous if and only if it is sequentially continuous.
  9. ^ a b A null sequence is a sequence that converges to the origin.
  10. ^ a b A sequence   is said to be Mackey convergent to 0 in   if there exists a divergent sequence   of positive real number such that   is a bounded set in  
  11. ^ If   is also a directed set under the usual function comparison then we can take the finite collection to consist of a single element.
  12. ^ In functional analysis, the strong dual topology is often the "standard" or "default" topology placed on the continuous dual space   where if X is a normed space then this strong dual topology is the same as the usual norm-induced topology on  
  13. ^ Technically, the topology must be coarser than the strong dual topology and also simultaneously be finer that the weak* topology.
  14. ^ Recall that a linear map is bounded if and only if it maps null sequences to bounded sequences.
  15. ^ For more information on such class of functions, see the entry on locally integrable functions.

References edit

  1. ^ a b Trèves 2006, pp. 222–223.
  2. ^ a b See for example Grubb 2009, p. 14.
  3. ^ a b Trèves 2006, pp. 85–89.
  4. ^ a b Trèves 2006, pp. 142–149.
  5. ^ Trèves 2006, pp. 356–358.
  6. ^ a b Rudin 1991, pp. 149–181.
  7. ^ a b c d Trèves 2006, pp. 131–134.
  8. ^ a b c d e f g h Trèves 2006, pp. 245–247.
  9. ^ a b Trèves 2006, pp. 247–252.
  10. ^ Trèves 2006, pp. 126–134.
  11. ^ a b c Trèves 2006, pp. 136–148.
  12. ^ Rudin 1991, pp. 149–155.
  13. ^ Narici & Beckenstein 2011, pp. 446–447.
  14. ^ a b Trèves 2006, p. 423.
  15. ^ See for example Schaefer & Wolff 1999, p. 173.
  16. ^ a b c d e Gabriyelyan, Saak "Topological properties of Strict LF-spaces and strong duals of Montel Strict LF-spaces" (2017)
  17. ^ a b Gabriyelyan, S.S. Kakol J., and·Leiderman, A. "The strong Pitkeev property for topological groups and topological vector spaces"
  18. ^ Trèves 2006, pp. 195–201.
  19. ^ Narici & Beckenstein 2011, p. 435.
  20. ^ Trèves 2006, pp. 526–534.
  21. ^ Trèves 2006, p. 357.
  22. ^ "Topological vector space". Encyclopedia of Mathematics. Retrieved September 6, 2020. It is a Montel space, hence paracompact, and so normal.
  23. ^ a b T. Shirai, Sur les Topologies des Espaces de L. Schwartz, Proc. Japan Acad. 35 (1959), 31-36.
  24. ^ According to Gel'fand & Shilov 1966–1968, v. 1, §1.2
  25. ^ Trèves 2006, pp. 351–359.
  26. ^ Narici & Beckenstein 2011, pp. 441–457.
  27. ^ a b Trèves 2006, pp. 150–160.
  28. ^ a b c d Trèves 2006, pp. 300–304.
  29. ^ Strichartz 1994, §2.3; Trèves 2006.
  30. ^ a b c Trèves 2006, pp. 131–135.
  31. ^ a b c Trèves 2006, pp. 240–245.
  32. ^ Trèves 2006, pp. 240–252.
  33. ^ Trèves 2006, pp. 262–264.
  34. ^ Trèves 2006, p. 218.
  35. ^ a b c Trèves 2006, pp. 255–257.
  36. ^ a b c d e f Trèves 2006, pp. 258–264.
  37. ^ Rudin 1991, pp. 177–181.
  38. ^ a b Trèves 2006, pp. 92–94.
  39. ^ Trèves 2006, pp. 160.
  40. ^ a b Trèves 2006, p. 531.
  41. ^ a b c d Trèves 2006, pp. 416–419.
  42. ^ Trèves 2006, pp. 509–510.

Bibliography edit

  • Barros-Neto, José (1973). An Introduction to the Theory of Distributions. New York, NY: Dekker.
  • Benedetto, J.J. (1997), Harmonic Analysis and Applications, CRC Press.
  • Folland, G.B. (1989). Harmonic Analysis in Phase Space. Princeton, NJ: Princeton University Press.
  • Friedlander, F.G.; Joshi, M.S. (1998). Introduction to the Theory of Distributions. Cambridge, UK: Cambridge University Press..
  • Gårding, L. (1997), Some Points of Analysis and their History, American Mathematical Society.
  • Gel'fand, I.M.; Shilov, G.E. (1966–1968), Generalized functions, vol. 1–5, Academic Press.
  • Grubb, G. (2009), Distributions and Operators, Springer.
  • Hörmander, L. (1983), The analysis of linear partial differential operators I, Grundl. Math. Wissenschaft., vol. 256, Springer, doi:10.1007/978-3-642-96750-4, ISBN 3-540-12104-8, MR 0717035.
  • Horváth, John (1966). Topological Vector Spaces and Distributions. Addison-Wesley series in mathematics. Vol. 1. Reading, MA: Addison-Wesley Publishing Company. ISBN 978-0201029857.
  • Kolmogorov, Andrey; Fomin, Sergei V. (1957). Elements of the Theory of Functions and Functional Analysis. Dover Books on Mathematics. New York: Dover Books. ISBN 978-1-61427-304-2. OCLC 912495626.
  • Narici, Lawrence; Beckenstein, Edward (2011). Topological Vector Spaces. Pure and applied mathematics (Second ed.). Boca Raton, FL: CRC Press. ISBN 978-1584888666. OCLC 144216834.
  • Petersen, Bent E. (1983). Introduction to the Fourier Transform and Pseudo-Differential Operators. Boston, MA: Pitman Publishing..
  • Rudin, Walter (1991). Functional Analysis. International Series in Pure and Applied Mathematics. Vol. 8 (Second ed.). New York, NY: McGraw-Hill Science/Engineering/Math. ISBN 978-0-07-054236-5. OCLC 21163277.
  • Schaefer, Helmut H.; Wolff, Manfred P. (1999). Topological Vector Spaces. GTM. Vol. 8 (Second ed.). New York, NY: Springer New York Imprint Springer. ISBN 978-1-4612-7155-0. OCLC 840278135.
  • Schwartz, Laurent (1954), "Sur l'impossibilité de la multiplications des distributions", C. R. Acad. Sci. Paris, 239: 847–848.
  • Schwartz, Laurent (1951), Théorie des distributions, vol. 1–2, Hermann.
  • Sobolev, S.L. (1936), "Méthode nouvelle à résoudre le problème de Cauchy pour les équations linéaires hyperboliques normales", Mat. Sbornik, 1: 39–72.
  • Stein, Elias; Weiss, Guido (1971), Introduction to Fourier Analysis on Euclidean Spaces, Princeton University Press, ISBN 0-691-08078-X.
  • Strichartz, R. (1994), A Guide to Distribution Theory and Fourier Transforms, CRC Press, ISBN 0-8493-8273-4.
  • Trèves, François (2006) [1967]. Topological Vector Spaces, Distributions and Kernels. Mineola, N.Y.: Dover Publications. ISBN 978-0-486-45352-1. OCLC 853623322.
  • Woodward, P.M. (1953). Probability and Information Theory with Applications to Radar. Oxford, UK: Pergamon Press.

Further reading edit