2023
[1]\fnmShiqing \surZhang
[1]\orgdivCollege of Mathematics, \orgnameSichuan University, \orgaddress\streetSouth Section 1 of the 1st Ring Road, \cityChengdu, \postcode610064, \stateSichuan, \countryChina
[2]\orgdivDepartment of Mathematics, \orgnameUniversity of Pennsylvania, \orgaddress\street209 South 33rd St, \cityPhiladelphia, \postcode19104, \statePennsylvania, \countryUSA
Weak Compactness Criterion in with an Existence Theorem of Minimizers
Abstract
There is a rich theory of existence theorems for minimizers over reflexive Sobolev spaces (ex. Eberlein-Šmulian theorem). However, the existence theorems for many variational problems over non-reflexive Sobolev spaces remain underexplored. In this paper, we investigate various examples of functionals over non-reflexive Sobolev spaces. To do this, we prove a weak compactness criterion in that generalizes the Dunford-Pettis theorem, which asserts that relatively weakly compact subsets of coincide with equi-integrable families. As a corollary, we also extend an existence theorem of minimizers from reflexive Sobolev spaces to non-reflexive ones. This work is also benefited and streamlined by various concepts in category theory.
keywords:
Dunford-Pettis theorem, non-reflexive Sobolev space, weak compactness in , existence theorem of a minimizer, calculus of variations.pacs:
[MSC Classification]46N20, 46E35, 28A20, 46E30
1 Introduction and Main Results
Historically, starting from Leonhard Euler and Joseph-Louis Lagrange, a series of practical problems in natural science including Fermat’s principle, the brachistochrone problem, and Dirichlet’s principle (see ZhangShiqing, ; Brezis, ; Courant, ; Courant&Hilbert, ), can be transformed into minimizing a functional of the following form:
where is a measurable space that satisfies some smoothness conditions, and varies in a suitable function space, is uniformly integrable on and known as the Lagrangian state function.
The methods for solving the above problems are collectively called the calculus of variations. These minimizers were initially considered to exist “naturally” until Karl Weierstrass constructed a counterexample (see Dacorogna, , Example 4.6 on page 122) in 1870. Since then, people have studied and obtained many existence theorems of minimizers. For example, Dirichlet’s principle was first proved in 1899 (with some strong conditions) then later extended to more general cases by David Hilbert Hilbert in 1904.
To be more precise, a real-valued functional induces a preorder over its domain such that for all in a Banach space :
A net equipped with the preorder above is called a decreasing net of . And whether a minimizer exists is equivalent to whether there is a greatest element for each of these decreasing nets. Therefore we should pay attention to whether any decreasing sequence satisfies that
(which is so called the minimizing sequence of ) can reach the infimum. Considering that the real line has a natural topology and selecting “what are the continuous functions from to ”, the topology on can be induced to (e.g. the norm topology and the weak topology) while keeping lower semicontinuous under this topology. Since lower semicontinuous functions can reach the infimum on compact sets, if a minimizing sequence is contained in some compact set, then there exists some in the closure of such that the functional can reach the infimum at .
In fact, the existence of minimizers depends on what domain space one chooses. Roughly speaking, the “larger” the space, the weaker the topology, the “fewer” open sets, and the easier it is for a set to be relatively compact. For aesthetic and practical reasons, one may typically expect a variational problem to have a solution as smooth as possible. However, it is usually difficult to directly find a solution with high regularity in a space with a strong topological structure. But if (weak) solutions can exist in a “larger” space, one only needs to verify that they are regular. For regularity, people have also developed many profound theories (see Evans, , Section 8.3 on page 458) and (Brezis, , Section 7.3 on page 191 and Section 9.6 on page 298), but we will only focus on existence. Arguably, Sobolev spaces (see Adams, , 3.2 on page 59) are tailor-made for this approach. By replacing classical derivatives with weak generalized derivatives, Sobolev spaces are “large” enough. But they are not too “bad” since even the smooth function space can be densely embedded in them (the Meyers-Serrin theorem Meyers ).
Assuming a functional is (weakly) lower semicontinuous in some Sobolev space, one would expect its corresponding minimizing sequence to be (weakly) relatively compact. For reflexive Sobolev spaces, we can apply the Eberlein-Šmulian theorem (see Yosida, , page 141), that is, a Banach space is reflexive if and only if any bounded set inside is weakly relatively compact. Based on this and James’s theorem James , Tang, Zhang, and Guo Tangyan have made a series of discussions on the relationship between the reflexivity of a Banach space and the existence of minimizers of a sequentially weakly lower semicontinuous functional on it. Using the Eberlein-Šmulian theorem, researchers have obtained some existence theorems of functional minimizers on reflexive Sobolev spaces. An example is Theorem 3.30 in (Dacorogna, , page 106), which can be used to prove Dirichlet’s principle as an application.
However, there are still various important practical problems whose integral functionals are defined on non-reflexive Sobolev spaces, such as the Plateau problem (see Courant, ; CourantR, ) (which is to show the existence of a minimal surface with a given boundary). Another simpler example follows:
Example 1.
(Dacorogna, , Example 4.5 on page 122 and Example 4.9 on page 124) Let vary in the following function space
where the value of on the boundary is defined in the sense of the trace operator. Let and be the functionals
and
Can their respective infimum be attained in ?
Since the Eberlein-Šmulian theorem no longer applies in this case, we need something stronger than boundedness to regain relative weak compactness. For the specific function space , Nelson Dunford and Billy James Pettis obtained the following theorem in 1940.
Theorem 1 (Dunford-Pettis theorem: weak compactness criterion in ).
(Dunford2, , Theorem 3.2.1 on page 376), (see also Brezis, , Problem 23 on page 466) Suppose that
- (hypothesis):
-
is a -finite measure space, and is a subset of .
Then the following statements and are equivalent. Here, the two statements have the following meaning.
- 1. :
-
is an equi-integrable family.
- 2. (weak compactness in ):
-
the set is relatively weakly compact in .
By an “equi-integrable family” in Theorem 1, we mean the following.
Definition 1 (equi-integrable family).
(Brezis, , 4.36 on page 129) A subset is said to be equi-integrable if it satisfies the following three conditions:
- a. (boundedness in ):
-
there exists such that for every .
- b. (equi-integrability):
-
there exists a function such that
for each , for any , and for all which is measurable in with its measure .
- c. (equi-integrability at infinity):
-
there exists a set-valued mapping from to (the power set of ) such that is measurable with its measure being finite, and there satisfies that
for every and any .
The notation here and thereafter refers to the Lebesgue measures (Adams, , page 14) in real Euclidean spaces. Note that when is bounded, the third condition is naturally true by choosing .
Remark 1.
We mark the hypothesis and statements in the Dunford-Pettis theorem with symbols for two reasons:
- 1.
-
We do not have to repeat the sentence from beginning to end every time. For example, we can shrink the Dunford-Pettis theorem to the following form:
We also use superscripts and subscripts to express some key information. For the full list of the types of symbols that will show up in this paper and their meaning, please see Appendix C.
- 2.
-
We can put the elements necessary for a sentence to have a specific meaning into the following brackets and emphasize them. For example, in , we know that if we fix and and can prove the existence of , then is a certain sentence. In fact, we can construct a thin category whose objects are sentences and morphisms are derivation symbols “”. It does not matter if we directly treat “” as “deduces” and “” as “is equivalent to”. Obviously “” are the isomorphisms in this category. See Appendix B for more details.
Based on the Dunford-Pettis theorem above, we will prove the following theorem on the non-reflexive Sobolev space through a technique of “disassembly and assembly”:
Theorem 2 (weak compactness criterion in ).
Suppose that
- (hypothesis):
-
is a non-empty open set in , and is a subset of .
Then the two statements and are equivalent. Here, the two statements have the following meaning.
- 1. :
-
is said to be equi-integrable under the sense of provided the following three conditions:
- a. (boundedness in ):
-
there exists a constant such that for every .
- b. (equi-integrability):
-
there exists a function such that
for every , for any , and for all which is measurable in with its measure . Here and henceforth, denotes a multi-index (Notation 1).
- c. (equi-integrability at infinity):
-
there exists a set-valued mapping such that is measurable with its measure being finite, and there satisfies that
for every and any .
- 2. (weak compactness in ):
-
is relatively weakly compact in .
When , Theorem 2 is exactly the Dunford-Pettis theorem if one regards the notation as .
When is bounded, the statement in Theorem 2 is always satisfied by taking . Moreover, if satisfies the cone condition (see Adams, , Paragraph 4.6 on page 82), we have the following theorem, which can also be seen as a high-dimensional generalization of Proposition 3.1.4 in Fathi .
Theorem 3.
Suppose that
- (hypothesis):
-
is a bounded open set in , and satisfies the cone condition. is a subset of .
Then the statement in Theorem 2 is equivalent to the following statement .
- :
-
there exists a constant such that
for every .
The statement in Theorem 2 can be replaced by the following statement .
- :
-
there exists a function such that
for every , for any , and for all which is measurable in with its measure .
Excitingly, this “generalized model” can also be “translated” to other well-known theorems. For example, we will generalize the Kolmogorov-M.Riesz-Fréchet theorem (see Brezis, , Corollary 4.27 on page 113) to obtain the following Corollary 4 and Corollary 5 as follows. After this, we will generalize the Ascoli-Arzelà theorem (see Brezis, , Theorem 4.25 on page 111) into Corollary 6 and Corollary 7. See Appendix A for the proofs of these corollaries.
Corollary 4 (compactness criterion in ).
Suppose that
- (hypothesis):
-
is a subset of with .
Then the statements and are equivalent. Here, the two statements have the following meaning.
- 1. :
-
the following three are satisfied:
- a. (boundedness in ):
-
there exists a constant such that for every .
- b. :
-
there exists a function such that
for every , for any , and for all with its norm . Here, the notation is a shift of the original function by the vector .
- c. :
-
there exists a set-valued mapping such that is measurable with its measure being finite, and there satisfies that
for every and any .
- 2. (compactness in ):
-
the set is precompact in .
Corollary 5.
Suppose that
- :
-
is a bounded open set in , and is a subset of with .
Then the statement in Corollary 4 is equivalent to as follows.
- :
-
there exists a constant such that
for every .
The statement in Corollary 4 can be replaced by the following .
- :
-
there exists a function such that
for every , for any , and for all with its norm . Here, the functions are extended to be outside .
Corollary 6 (compactness criterion in ).
Suppose that
- (hypothesis):
-
is a compact metric space which satisfies the uniform regularity condition (see Adams, , 4.10 on page 84), and is a subset of .
Then the two statements and are equivalent. Here, the two statements have the following meaning.
- 1. :
-
is said to be uniformly equicontinuous under the sense of provided the following two conditions:
- a. (boundedness in ):
-
there exists a constant such that for every .
- b. (equicontinuity):
-
there exists a function such that
for every , for any , and for all and in with the distance between them .
- 2. (compactness in ):
-
the set is precompact in .
Corollary 7.
Statement in Corollary 6 can be replaced by the following .
- :
-
there exists a function such that
for every , for any , and for all and in with the distance between them .
Remark 2.
For an arbitrary satisfying in Corollary 6, we have not been able to generalize to only verify the boundedness of when is or . We considered that continuous functions are very similar to Lebesgue space , and on finite measure space can be regarded as the intersection of all spaces with . For each , there exists a constant such that
for every . However might be .
When is a subset of , the Landau–Kolmogorov inequality Landau can be considered to apply. But our topic is not interpolation inequality so we will not go into it too much.
With the compactness criterion in , we can extend some theorems that were previously proven in reflexive spaces to some non-reflexive spaces. As an application, we will prove the following theorem as an extension of Theorem 3.30 in (Dacorogna, , page 106):
Theorem 8 (an existence theorem of minimizers in ).
Let be a bounded open subset of , and satisfies the cone condition. Let be a Carathéodory function (see Dacorogna, , Definition 3.5 on page 75) satisfying
(1) |
for almost every and for every and for some , . And let the functional
be finite at some . Then every minimizing sequence in is bounded.
In addition, if is convex, and a minimizing sequence is found to satisfy the statement . Then attains its minimum at some .
Furthermore, if is strictly convex for almost every , the minimizer is unique.
Applying Theorem 2, we can also get the following proposition:
Proposition 9.
Let the functional be defined on
with the form
for some , , and . Then the infimum of can be achieved at some if and only if .
Moreover, if the infimum is attainable, the derivative is non-negative for almost every .
Applying Proposition 9 we can directly answer the question in Example 1, that is, cannot reach its infimum, but can. In fact, has infinitely many minimum points.
Outline The rest of the paper is organized as follows. In Section 2, we introduce two important definitions and prove a technical lemma (Lemma 11) that will be instrumental in the proof of Theorem 2. In Section 3, we prove Theorem 2, Theorem 8, and Proposition 9. In Appendix A, we prove Corollary 4 and Corollary 6. In Appendix B, we introduce more details about the syntactic category. In Appendix C, we provide a collection of notations and abbreviations used in this paper.
2 Definitions and Lemmas
To prove Theorem 2, we introduce an operator (see Definition 3) that embeds the Sobolev spaces into some spaces. When considering a Sobolev space defined on a non-empty open subset in a real Euclidean space , we assign a replica that is “exactly the same” as the open set to each multi-index , and take the (disjoint) union of these replicas (see Definition 2) as the domain of the functions in the space . We prove that the operator is an isometry (see Lemma 10). By converting it into a problem about the space , we can apply Theorem 1 to complete the proof of Theorem 2.
Definition 2 (Construction of the disjoint union ).
(Adams, , 3.5 on page 61) Let be a non-empty open set in . For each , let be a different copy of lying in which is a different copy of with respect to the multi-index . Thus these non-empty open sets are mutually disjoint. More mathematically, an isomorphism between and refers to an isometry that also preserves the measure:
Here, denotes the Lebesgue measure in .
The union of these sets is denoted as
The space naturally inherits the following structures from the Euclidean spaces:
- Measure:
-
A subset is measurable if is Lebesgue measurable in for all . The measure of such in is defined as the sum of the Lebesgue measures of taken over , i.e.
- Topology:
-
The space is equipped with the disjoint union topology, i.e., a set is open in if and only if is open in for every .
Now we can define the isometry operator from the Sobolev space to the Lebesgue space . In this way, we transform the discussion on the compactness of a subset of into the discussion on the compactness of its image in space .
Definition 3 (Construction of the isometric operator ).
(Adams, , 3.5 on page 61) We construct an operator from to as follows:
where is given by
That is, the restriction of function on each component of in is given by the function .
Remark 3.
From the perspective of category theory, let the objects in an indicator category be the multi-indies, and the morphisms are only the identities. Then is the colimit of a which maps each to . And is the limit of a (contravariant) functor which maps each to . Therefore the operator is the unique morphism from to . See Chapter 1-3 of riehl2017category for a precise explanation of these terminologies, and see the two illustrations in Figure 1 and Figure 2.
Lemma 10.
The operator is a well-defined isometry. Let be the range of . is a closed linear subspace of . Furthermore, and are homeomorphic both in their norm and weak topologies.
Proof: [Proof of Lemma 10] We first check that is well-defined. Notice that for any , there must be only one multi-index such that . Furthermore, for each , the weak generalized partial derivative is unique up to sets of measure zero in . Thus is well-defined.
Now we will check that is an isometry. Indeed, we have the following equalities
Since is an isometry between two complete metric spaces, the image of is also complete and thus closed in . Finally, it follows from (see Brezis, , Theorem 3.10 on page 61) that and are homeomorphic both in their norm topologies and their weak topologies.
Following this lemma, we denote the inverse operator of as . As a useful result, we have the following lemma.
Lemma 11.
Suppose that (the same as Theorem 2)
- (hypothesis):
-
is a non-empty open set in , and is a subset of .
Then the following statements are equivalent:
- (weak compactness in ):
-
is relatively weakly compact in .
- :
-
is relatively weakly compact in .
Proof: [Proof of Lemma 11.]
We first need to prove that under the hypothesis , the following two topological spaces are homeomorphic.
- a.
-
: the Sobolev space equipped with the weak topology.
- b.
-
: the space equipped with the subspace topology induced from the weak topology of the space , where .
Recall that the basis elements for the topology can be written as
(2) |
for some finite , and , , .
Using the analytic form of Hahn-Banach theorem (see Brezis, , Theorem 1.1 on page 1), we have that for every , there exists such that
Thus for each in (2), there exists such that
It means that is stronger than .
Besides, similarly, every open set for is a union of such sets:
(3) |
with being a finite set, , , .
In summary, we obtain that and have the same topological basis. Thus
We know from Lemma 10 that is a closed linear subspace. Moreover, is weakly closed by its convexity (linearity) using Mazur’s lemma (see Yosida, , Theorem 2 on page 120). It follows that relatively compact sets for coincide with relatively weakly compact sets in . Hence is relatively weakly compact in if and only if is relatively weakly compact in .
3 Proofs of Main Results
In this section we prove the main results in Section 1. We leave the proof of Corollary 4, 5, 6, and 7 to Appendix A.
Proof: [Proof of Theorem 2.] The idea is to show the following list of equivalences:
(4) | ||||
(5) | ||||
(6) | ||||
(7) | ||||
(8) |
For convenience of the reader, we summarize the main idea behind proving the equivalences of these five statements in Figure 3.
Now we move on to the actual proof. As a result that will be used repeatedly, we first prove that:
(9) |
In fact, since is a -finite measure space (as an open set in which is isometric with ), we have for every . By the definition of the general product in (see Definition 4), we immediately obtain that:
Moreover, since is also a -finite measure space, we have
Thus we get (9). Herein we show (4) (5) (8) one by one. These are all simple verifications.
Suppose , and let
It is not difficult to verify that:
-
1.
For any , its norm satisfies
which implies for every .
-
2.
Similarly, for each , for any and for every measurable set with its measure , since , we have
Thus we get for every .
-
3.
Again, for each and for any , the inequality
implies for each .
Hence by the definition of general products, there exists the morphism “(4) (5)”.
Suppose , and let
We obtain that:
-
1.
For any , its norm satisfies that
which implies .
-
2.
For each , for any and for every measurable set with its measure , since for all , we have
Thus we get .
-
3.
In the same way, for each and any , the inequality
implies .
Suppose , and let
Almost the same as the previous argument, we can verify that:
-
1.
For every , its norm satisfies
which implies .
-
2.
For each , for any , and for every measurable set with its measure , since for all , we have
Thus we get .
-
3.
For each and any , we have the inequality
which leads to immediately.
Suppose , and let
We obtain that:
-
1.
For every , its norm satisfies
which implies for every .
-
2.
For each , for any , and for every measurable set with its measure , we have
Thus we get for every .
-
3.
For each , the inequality
implies for every .
Isomorphism “(6) (7)” can be obtained immediately from Theorem 1. Using Lemma 11 and (9), isomorphism “(7) (8)” is obvious.
Finally, the proof of Theorem 2 is completed by compositing.
Proof: [Proof of Theorem 3.] We only need to prove the following two.
- 1.
-
.
It can be obtained directly from the interpolation inequality (see Adams, , Theorem 5.2 on page 135).
- 2.
-
.
From the above discussion we already know that is bounded in . Applying the Rellich-Kondrachov theorem (see Adams, , Theorem 6.3 on page 168), we obtain that the embedding from to is compact (When , the notation is regarded as ). Hence is precompact in and obviously, it is relatively weakly compact. By Lemma 11, we have that is relatively weakly compact in . Then applying Theorem 2, the statement is satisfied.
Thus the proof is completed.
Proof: [Proof of Theorem 8.] Denote
and observe that . Integrating both sides of (1) with respect to , we have
(10) |
Next, we show that . Let for some . Invoking Poincaré’s inequality (see Brezis, , Corollary 9.19 on page 290), there exists a positive constant such that
Combining with the triangle inequality, we have
In short,
(11) |
Denote
According to (10) and (11), we obtain that
(12) |
Thus we have proved that
Hence, for an arbitrary minimizing sequence and then for some integer sufficiently large, we obtain that
Combining with (12), we have
It means that there exists a number such that
If we could find some suitable sequence turning out to satisfies the statement in Theorem 3, we should have that is relatively weakly compact. Thus there exists a subsequence of (also denoted as for simplicity) converging weakly to some .
Using Theorem 3.23 in (Dacorogna, , page 96) to obtain that functional is weakly lower semicontinuous, we have
which implies that attained its infimum at .
It remains to show the uniqueness of under the condition that the function is strictly convex. Here we give the proof by contradiction.
Assume that there exists different from but . By the strict convexity, for every , we have that
(13) |
Integrating both sides of the above inequality with respect to , we obtain
Since the integrand is non-negative, the inequality above becomes an equality if and only if the integrand is almost everywhere, which is in contradiction to (13). Hence we have
However, it contradicts the fact that is the infimum. Thus, the uniqueness part has been proved. This completes the proof of the theorem.
Proof: [Proof of Proposition 9.] First, let us consider a simpler case. Suppose that there exist some and such that
We have
Take . Then we have that
For any minimizing sequence , using Theorem 8, is bounded. Denote
Obviously that and .
If , we have that . Then is also a minimizing sequence in . Since is bounded, we obtain that and
And it implies that . For each , denote
Since , the Lebesgue measure . But , which contradicts Statement in Theorem 3. Hence the minimizing sequence is not relatively weakly compact in .
If , using the Newton-Leibniz formula, there holds the inequality
with equality if and only if for almost everywhere in . Obviously, every monotonically increasing function with respect to minimizes the functional already.
In conclusion, the functional
has a minimal point if and only if . Moreover, we also obtain that and for almost everywhere in .
Next, we consider the case where the integrand has the form
Thus we have
If , let and .
The following inequality will be used later:
(14) |
with equality if and only if , where , and .
We proceed to prove (14). Obviously, if , the equal sign holds, so does (14). Besides, if , denote . Notice that and , thus
Hence the proof of (14) is completed.
Take and in (14). We have
(15) |
Let
be two functionals defined on . According to the discussion above, the infimum of is and can be achieved. Since every sequence minimizing satisfies that , we have
Let be a minimizing sequence of satisfying and . According to (16), . It follows that
Therefore, is necessary for to attain the infimum. It remains to show that is sufficient too.
If , we have
Using the Newton-Leibniz formula, for every satisfying that for almost everywhere in , there holds . Thus, the sufficiency is proved.
Finally, since is a constant, it does not affect whether the infimum of the functional can be achieved. This completes the proof of Proposition 9.
Acknowledgments
This work was supported partially by NSF of China, No. 12071316. We would also like to thank Professor Zhu Chaofeng from Nankai University for his discussion on whether interpolation inequalities in continuous differentiable function spaces can be applied. We also thank Professor Zhou Feng from East China Normal University for his fruitful comments about the boundary conditions of the domain in each theorem. M.J. would like to thank Professor Hongjie Dong from Brown University for helpful conversation.
Data availability statement This manuscript has no associated data.
Statements and Declarations
Conflict of interest The authors declared that they have no conflicts of interest to this work.
Appendix A Proofs of Corollaries
Proof: [Proof of Corollary 4.]
The proof of this corollary is almost “isomorphic” to the proof of Theorem 2.
Proof of “”:
Of course, we suppose that . Then
-
1.
For any , its norm satisfies
-
2.
For any , and with , we have
-
3.
Take . For any and , we have
Using the Kolmogorov-M.Riesz-Fréchet theorem (see Brezis, , Corollary 4.27 on page 113), we obtain that “ is precompact in ”, which is equivalent to that “ is precompact in ”.
Proof of “”:
Similarly, suppose we are given . By the definition of , we have that is precompact in . Using the Kolmogorov-M.Riesz-Fréchet theorem again, we obtain that
-
1.
There exists a positive constant such that for any , its norm satisfies . Thus we have
which implies .
-
2.
There exists a function such that for any , and with , there holds . Then we have
Thus we get .
-
3.
There exists a set-valued mapping such that for every and , there holds . Take , then we have
which leads to immediately.
The proof is completed.
Proof: [Proof of Corollary 5.] The proof of this corollary is almost “isomorphic” to the proof of Theorem 3. Similarly, we only need to prove the following two.
- 1.
-
.
It can be obtained directly from the interpolation inequality (see Adams, , Theorem 5.2 on page 135).
- 2.
-
.
From the above discussion we already know that is bounded in . Applying the Rellich-Kondrachov theorem (see Adams, , Theorem 6.3 on page 168), we obtain that the embedding from to is compact (When , the notation is regarded as ). Hence is precompact in . Then applying Corollary 4, the statement is satisfied.
Thus the proof is completed.
Proof: [Proof of Corollary 6.] We need to fine-tune the definition of the isometric operator so that it ”translates” from the Sobolev space to the space of continuously differentiable functions. Denote
where is a disjoint copy of given by the isometric mapping . Moreover, suppose satisfies that for each fixed there holds
i.e., the distance between two different “”s is greater than the diameter of but less than , which is always easy to satisfy. By these assumptions, we have that is a compact metric space. Then we denote
where
It is easy to verify that is well-defined. Let be the range of , so is a closed subspace of . The same arguments as in Lemma 10 remind us is an isometry from to . We denote its inverse operator as .
Proof of “”:
Suppose . Then
-
1.
For any , its norm satisfies
-
2.
Take . For any , and with , since and must come from a same “” (without loss of generality, we name it ), thus we have
Using the Ascoli-Arzelà theorem (see Brezis, , Theorem 4.25 on page 111), we obtain that “ is precompact in ”, which is equivalent to that “ is precompact in ”.
Proof of “”:
Suppose that we are given . By the definition of , we have that is precompact in . Using the Ascoli-Arzelà theorem again, we obtain that
-
1.
There exists a positive constant such that for any , its norm satisfies . Thus we have
which implies .
-
2.
There exists a function such that for any , and with (note that ), there holds for every . Then we have
Thus we get .
The proof is completed.
Proof: [Proof of Corollary 7.] We only need to show
-
.
Note that the embedding from to is compact. It follows that the embedding from to is compact. Hence is precompact in . Then applying Corollary 6, the statement is satisfied.
The proof is completed.
Appendix B Syntactic Category
To describe and prove Theorem 2 more clearly, we introduce a few notations from category theory, but we will not be using some overly abstract concepts. Inspired by the syntactic category (see John, , page 10), we simplify sentences into symbols and make the following definition, which is different from the original meaning.
Definition 4.
Let be a set of sentences in some propositional language. For simplicity in this paper, we encourage the reader to think of as “a collection of mathematical statements with some axioms”. We define the synatic category determined by as the category with the following data:
- Objects:
-
The sentences in and those formed by connecting some elements in with the connectives any number of times. In particular, in this paper, all sentences represented by abbreviations are elements of . For example, , , and (in Theorem 1) are all in .
- Morphisms:
-
For objects and , the morphism coincides with a proof from to . For example, the Bolzano-Weierstrass (BW) theorem gives a morphism “” from the sentence “” to the sentence “”.
We assume that each sentence can be proved from itself without any reason, that is, for every . Since we only focus on the existence of proofs, we will omit the information about what the proof is actually on the arrows, that is, we treat different proofs e.g. and as a same morphism hereafter. In other words, contains at most one element for every and . It also means that is a thin category.
- Composites:
-
given two morphisms and , the composite is naturally the proof obtained by concatenating the two proofs.
Note that unlike the original definition (see John, , page 10), we do not require that the sentences of the proofs must be in , that is, we can use any mathematical theorem in a self-consistent axiom system as a morphism. Therefore, what elements the set contains does not affect whether there exists a morphism between two objects in . The following categorical data may be interpreted in as follows:
- Isomorphisms:
-
an isomorphism is a morphism such that there exists a morphism as the inverse. We denote isomorphisms by .
- General products:
-
(see Abramsky, , page 22) for a family of objects , a product is an object with canonical proofs of conjuncts from conjunctions as projection morphisms for each . Products exist in universally.
- Uniqueness up to unique isomorphism:
-
(see Abramsky, , page 23) suppose that and are both general products of objects , then and are isomorphic, i.e., .
Remark 4.
Dually, one may define coproducts using the connectives (i.e. or).
Remark 5.
When we add a new sentence into such that the category expands to , one may observe that is a full subcategory of , therefore morphisms in are inherited into . We can always make contain enough sentences, so we do not strictly define what contains hereafter.
Appendix C Notations and Abbreviations
The following notations and abbreviations are used in this manuscript:
The Lebesgue measure in real Euclidean spaces | |
And | |
Is equivalent to | |
Deduces | |
The multi-index (see Notation 1) | |
The length of multi-index | |
The weak generalized partial derivative of | |
A copy of with respect to (see Definition 2) | |
An isometry from to | |
The Lebesgue measure in | |
The disjoint union of all the with | |
The measure in | |
A special isometry (see Definition 3) |
Sentences for hypotheses | |
is a -finite measure space, and is a subset of . | |
is a non-empty open set in , and is a subset of . | |
is a bounded open set in , and satisfies the cone condition. is a subset of . | |
is a subset of with . | |
is a compact metric space which satisfies the uniform regularity condition, and is a subset of . |
Statement related to boundedness | |
There exists such that for every . | |
There exists such that for every . | |
There exists a constant such that for every . | |
There exists a constant such that for every . |
Statement related to equi-continuity. | |
There exists a function such that for every , for any , and for all and in with the distance between them . | |
There exists a function such that for every , for any , and for all and in with the distance between them . |
Statement related to equi-integrability. | |
. | There exists a function such that for every , for any , and for all measurable set with its measure . |
There exists a function such that for every , for any , and for all which is measurable in with its measure . | |
There exists a function such that for every , for any , and for all which is measurable in with its measure . | |
There exists a function such that for every , for any , and for all with its norm . | |
There exists a function such that for every , for any , and for all with its norm . Here, the functions are extended to be outside . |
Sentences for some equi-properties at infinity | |
There exists a set-valued mapping such that is measurable with its measure being finite, and there satisfies that for every and any . | |
There exists a set-valued mapping such that is measurable with its measure being finite, and there satisfies that for every and any . |
Statement related to compactness | |
is relatively weakly compact in . | |
is relatively weakly compact in . | |
is precompact in . | |
is precompact in . |
Notation 1 (multi-indices ).
(Adams, , 3.5 on page 61) Given integers and , let
be a multi-index such that
For any where , we write
to denote the weak generalized partial derivative of with respect to .
References
- \bibcommenthead
- (1) Zhang, S.: Functional Analysis with Applications(in Chinese). Science Press, Beijing (2018)
- (2) Brezis, H.: Functional Analysis, Sobolev Spaces and Partial Differential Equations. Springer, New York (2011)
- (3) Courant, R.: Dirichlet’s Principle, Conformal Mapping, and Minimal Surfaces. Springer, New York-Heidelberg (1977)
- (4) Courant, R., Hilbert, D.: Methods of Mathematical Physics: Partial Differential Equations. John Wiley & Sons, New York-Chichester-Brisbane-Toronto-Singapore (1989)
- (5) Dacorogna, B.: Direct Methods in the Calculus of Variations. Springer, New York (2008)
- (6) Hilbert, D.: Über das Dirichletsche Prinzip. Math. Ann. 59, 161–186 (1904). https://doi.org/10.1007/BF01444753
- (7) Evans, L.C.: Partial Differential Equations. American Mathematical Society, Providence, RI (2010)
- (8) Adams, R.A., Fournier, J.J.F.: Sobolev Spaces. Elsevier/Academic Press, Amsterdam (2003)
- (9) Meyers, N.G., Serrin, J.: . Proc. Nat. Acad. Sci. U.S.A. 51, 1055–1056 (1964). https://doi.org/10.1073/pnas.51.6.1055
- (10) Yosida, K.: Functional Analysis. Springer, Berlin (1995)
- (11) James, R.C.: Characterizations of Reflexivity. Studia Mathematica 23, 205–216 (1964)
- (12) Tang, Y., Zhang, S., Guo, T.: Some Nonlinear Characterizations of Reflexive Banach Spaces. Journal of Mathematical Analysis and Applications 519, 126736 (2023). https://doi.org/10.1016/j.jmaa.2022.126736
- (13) Courant, R.: Plateau’s Problem and Dirichlet’s Principle. Ann. of Math. (2) 38, 679–724 (1937). https://doi.org/10.2307/1968610
- (14) Dunford, N., Pettis, B.J.: Linear Operations on Summable Functions. Trans. Amer. Math. Soc. 47, 323–392 (1940). https://doi.org/10.2307/1989960
- (15) Fathi, A.: Weak KAM Theorem in Lagrangian Dynamics Preliminary Version Number 10. by CUP (2008). Available at https://www.math.u-bordeaux.fr/~pthieull/Recherche/KamFaible/Publications/Fathi2008_01.pdf
- (16) Landau, E.: Einige Ungleichungen Fur Zweimal Differentiierbare Funktionen. Proceedings of the London Mathematical Society. Second Series 13, 43–49 (1914). https://doi.org/10.1112/plms/s2-13.1.43
- (17) Riehl, E.: Category Theory in Context. Dover Publications, New York (2017)
- (18) Bell, J.L.: The Development of Categorical Logic. In: Gabbay, D.M., Guenthner, F. (eds.) Handbook of Philosophical Logic, vol. 12, pp. 279–360. Springer, New York (2005)
- (19) Abramsky, S., Tzevelekos, N.: Introduction to Categories and Categorical Logic. In: Coecke, B. (ed.) New Structures for Physics, pp. 3–94. Springer, Heidelberg (2011)