Density estimate from below
in relation to a conjecture of A. Zygmund
on Lipschitz differentiation
Abstract.
Letting be Borel measurable and Lipschitzian, we establish that
for -almost every . In particular, it follows that is -negligible if and only if , for -almost every .
Key words and phrases:
Lebesgue measure, Nikodým set, Negligible set, Derivation basis, Zygmund conjecture, Lipschitz differentiation2010 Mathematics Subject Classification:
Primary 28A75,26B151. Foreword
Let be the set of squares centered at . The Lebesgue density theorem states that if is Lebesgue summable, then
() |
for -almost every . This consequence of the Vitali covering theorem fails if is replaced with , the set of rectangles centered at of arbitrary direction and eccentricity. It fails even for some indicator function . Indeed, Nikodým [12] defined a set of full measure with the following property: For every , there exists a line such that . In fact, Nikodým’s example shows that the Lebesgue density theorem fails if we replace with , the set of rectangles of arbitrary eccentricity, centered at , and one side of which is parallel to . Furthermore [4, Chap. IV Theorem 3.5], replacing by a nonnegligible subset of , one may choose to be continuous with respect to . Yet, if is constant, then the Lebesgue density theorem with respect to holds, for -summable functions , , by virtue of a theorem of Zygmund, though the corresponding version of the Vitali covering theorem fails, according to an example of H. Bohr [4, Chap. IV Theorem 1.1]. This raises the question: What regularity condition of guarantees that the Lebesgue density theorem holds with respect to for, say, functions that are square summable? The following version is a conjecture reportedly [11] attributed to Zygmund. Assume that is a Lipschitzian field of lines, with , and is square summable. Is it true that
for -almost every ? Here, is the 1-dimensional Hausdorff measure.
E.M. Stein raised the singular integral variant of this conjecture. Both have received much attention from the harmonic analysis community. To the author’s knowledge, most results recorded so far in the literature, via the maximal function approach, assume some extra regularity property of – namely, that be together with a hypothesis on the variation of the derivative – see, for instance, [1] and [11].
In this paper we offer a novel approach – a kind of change of variable based on an appropriate fibration and the coarea formula. This allows for treating the case when is Lipschitzian, without further restriction. We obtain the following lower density bound for indicator functions. Let be a Lipschitzian field of lines, with , and let be Lebesgue measurable. Then,
for -almost every . Incidentally, the following corollary – a nonparallel version of Fubini theorem – seems to be new as well. The set is Lebesgue negligible if and only if , for -almost every .
Our results hold in any dimension and codimension.
2. Sketch of proof
Let be a subset of Euclidean space , , and let be the Lebesgue outer measure. We start by considering the following weak question: Can one tell whether is Lebesgue negligible from the knowledge only of its trace on each member of some given collection of “lower dimensional” subsets , . One expects that if is “negligible in the dimension of ”, for each , then . Of course, a necessary condition is that the sets cover almost all of , i.e., . Consider, for instance, , , and , for , the collection of all vertical lines in the plane. It is not true in general that if and is a singleton, for each , then . There exist, indeed, functions whose graph satisfies – see, e.g., [10, Chapter 2 Theorem 4] for an example due to Sierpiński. In order to rule out such examples, we will henceforth assume that is Borel measurable. In that case, the theorem of Fubini, together with the invariance of the Lebesgue measure under orthogonal transformations, imply the following. Given an integer , if is the collection of all -dimensional affine subspaces of of some fixed direction, and if for all , then . Here, denotes the -dimensional Hausdorff measure. A special feature of this collection is that it partitions , its members being the level sets , , of a “nice map” , indeed, an orthogonal projection. This is an occurrence of the following more general situation when and its leaves are allowed to be nonlinear. The coarea formula due to Federer [8] asserts that if is Lipschitzian and is Borel measurable, then
Thus, if the Jacobian coarea factor is positive, -almost everywhere in , then the collection is suitable for detecting whether or not is Lebesgue null. At -almost all , the map is differentiable, by virtue of Rademacher’s theorem, and
In this paper, we focus on the case when , , are affine subspaces of , but not necessarily members of a partition of the ambient space. Specifically, we assume that with each is associated an -dimensional affine subspace of containing . Given a Borel set , the question whether
(1) |
has a negative answer, in view of Nikodým’s set evoked in the previous section. Indeed, corresponding to this set , there exists a field of lines such that , for all . In this context, a selection theorem due to von Neumann implies that (possibly considering a smaller, non Lebesgue null, Borel subset of ) the correspondence can be chosen to be Borel measurable (see 3.19) and, in turn, it can be chosen to be continuous, according to Lusin’s theorem. Nonetheless, when is Lipschitzian, the situation improves, as illustrated in our theorem below; is the Grassmannian manifold consisting of -dimensional vector subspaces of .
Theorem. —
Assume is Lipschitzian and is Borel measurable. The following are equivalent.
-
(A)
.
-
(B)
For almost every , .
-
(C)
For almost every , .
As should be apparent from the discussion above, one difficulty stands with the fact that the affine -planes may not be disjointed. Nevertheless, they locally are, in the following sense. Given there exist a neighborhood of and Lipschitzian maps so that is an orthonomal frame spanning , for , and the map is a lipeomorphism of a neighborhood of onto its image – here, and is close to . This, and applications of Fubini’s theorem, yield in the theorem above (see 8.2).
However, we aim at obtaining a quantative version of this result, that the change of variable just described does not seem to provide. A natural route is to reduce the problem to applying the coarea formula, by spreading out the ’s in a disjointed way, in a higher dimensional space – i.e., adding a variable to the given and considering as a fiber above the base space . We thus define
where is Borel measurable. This set is -rectifiable, owing to the Lipschitz continuity of . It is convenient to assume that , so that
(2) |
, is a locally finite Borel measure 3.16. Now, was precisely set up so that, for each ,
where and denote the projections of to the and variable, respectively. Abbreviating , the coarea formula yields
(3) |
A simple calculation 5.4 shows that , -almost everywhere on . Since also
(4) |
the implication above should now be clear: Letting and , one infers from hypothesis (A) and (4) that , whence, , by (3), and, in turn, conclusion (C) ensues from (2).
In order to establish that by using this new change of variable, we need to observe 5.2 that , -almost everywhere, and, ideally, to show that , for almost every . This last part involves some difficulty. To understand this, we let , in order to keep the notations short. Now iff iff , where is, say, a unit vector. Abbreviating , we infer that
The problem remains that two of the nonlinear sets and may intersect, thereby preventing another application of the coarea formula when looking out for their lower bound. Yet, we already know that
where is a Radon-Nikodým derivative (see 5.6) and also that is comparable to (see 5.8). Adding an extra variable to the fibered space , 5.10, we improve on this by showing that
where the last equality defines , and is a lower bound for the coarea Jacobian factor of the restriction to the fibered space of the projection (see 5.12 and 5.16). We are reduced to showing that , almost everywhere. The reason why this holds is the following. Fix a Borel set , and . Let be the cylindrical box consisting of those such that and . We want to find a lower bound for
To this end, we fix and we let denote the corresponding vertical line segment. According to Fubini’s theorem we are reduced to estimating
By virtue of Vitali’s covering theorem, we can find a disjointed family of line segments , covering almost all , such that the above integral nearly equals
where the first near equality follows from the coarea formula, and the second one because at small scales (see 3.12) and the “nonlinear horizontal stripes” are nearly pairwise disjoint. Verification of these claims takes up sections 6 and 7. Now, we reach a contradiction if is assumed to have , and is a point of density of .
Along this path, we have, in fact, achieved a quantitative version of these observations. Indeed, assuming that , for positively many , we start by choosing with the same property, whose diameter is small enough for the various estimates above to hold, and we refer to Egoroff’s theorem to select so that the set
has positive measure. Thus, for , there exists and there exists as small as we please so that
where the symbol means that we loose a multiplicative factor as close as we wish to 1, according to the scale . The above contradiction proves our main result 8.4, namely,
Theorem. —
Assume is Lipschitzian and is Borel measurable. Abbreviating , it follows that
for -almost every .
I extend my warm thanks to Jean-Christophe Léger for carefully reading an early version of the manuscript.
3. Preliminaries
3.1. —
In this paper, are integers. The ambient space is . The canonical inner product of is denoted and the corresponding Euclidean norm of is . If , we let be the -algebra of Borel measurable subsets of .
3.2Hausdorff measure. —
We let be the Lebesgue outer measure in and . For , we abbreviate . Given , we call -cover of a finite or countably infinite family of subsets of such that and , for every . We define
and . Thus, is the -dimensional Hausdorff outer measure in .
-
(1)
If is a sequence of nonempty compact subsets of converging in Hausdorff distance to , then .
Given , choose an -cover of such that . Since , for each , we can choose an open set containing such that . Since is open, there exists a positive integer such that , whenever . Thus, in that case is an -cover of and, therefore, . Take the limit superior of the left hand side, as , and then let .
-
(2)
For all , one has .
It suffices to note that . The first inequality is trivial; the second one follows from the isodiametric inequality [9, 2.10.33]; the last one is a consequence of the Vitali covering theorem [6, Chapter 2 §2 Theorem 2].
-
(3)
If is an -dimensional affine subspace and , then .
Let denote the -dimensional Hausdorff outer measure in the metric space . In other words,
and . It is elementary to observe that and that . Now, if is an isometry, then , where the third equality follows from claim (2) above.
3.3Coarea formula. —
We recall two versions of the coarea formula. First, if is -measurable and is Lipschitzian, then is - measurable and
Here, the coarea Jacobian factor is well-defined -almost everywhere, according to Rademacher’s theorem, and equals
see, for instance, [6, Chapter 3 §4].
Secondly, if is -measurable and countably -rectifiable and if is Lipschitzian, then is -measurable and
In order to state a formula for the coarea Jacobian factor of relative to , we consider a point , where admits an approximate -dimensional tangent space and is differentiable at along . Letting denote the derivative of at , we have
see, for instance, [9, 3.2.22].
In both cases, it is useful to recall the following. If is a linear map between two inner product spaces and , then
(5) |
On the one hand, [9, 1.7.6] and , with as above. On the other hand, if are linearly independent vectors of , then
(6) |
Finally, we observe that both coarea formulæ hold true when is merely locally Lipschitzian, according to the monotone convergence theorem.
3.4Grassmannian. —
We let be the set whose members are the -dimensional vector subspaces of . With , we associate , the orthogonal projection on . We give the structure of a compact metric space by letting , where is the operator norm. If , then satisfies , therefore, is an isometry. The bijective correspondence identifies with the submanifold . There exists an open neighborhood of in and a Lipschitzian retraction , according, for instance, to [9, 3.1.20]. Therefore, if and is Lipschitzian, then there exist an open neighborhood of in and a Lipschitzian extension of . Indeed, admits a Lipschitzian extension – see, e.g., [9, 2.10.43] – and it suffices to let and .
3.5Orthonormal frames. —
We let be the set of orthonormal -frames in , i.e., . We will consider it as a metric space, with its structure inherited from .
3.6. —
Let be a nonempty closed set such that . There exists a Lipschitzian map such that , for every .
Proof.
Pick arbitrarily . If , then the map is bijective: If , then , thus, . Letting be an arbitrary basis of , it follows that, for each , the vectors , , constitute a basis of . Furthermore, the maps are Lipschitzian: . We apply the Gram-Schmidt process:
so that is readily an orthogonal basis of depending upon in a Lipschitzian way. Since each is bounded away from zero on , the formula , , defines with the required property. ∎
3.7. —
There exists a Borel measurable map with the property that , for every .
Proof.
Since is compact, it can partitioned into finitely many Borel measurable sets , each having diameter bounded by . Define , piecewise, to coincide on with associated to in 3.6, . ∎
3.8. —
Assume , , and is Lipschitzian. There exist an open neighborhood of in and Lipschitzian maps with the following properties.
-
(1)
For every , the family is an orthonormal basis of .
-
(2)
For every , one has
and
Proof.
3.9. —
Assume is Borel measurable. There exist Borel measurable maps with the following properties.
-
(1)
For every , the family is an orthonormal basis of .
-
(2)
For every , one has
and
Proof.
Choose and as in 3.7. Letting and , for , completes the proof. ∎
3.10Definition of . —
The typical situation that arises in the remaining part of this paper is that we are given a set , a Lipschitzian map , and . We will represent and in a neighborhood of as in 3.8. We will then further reduce the size of several times, in order that various conditions be met. With no exception, we will denote as the affine subspace containing , of direction , whenever is defined.
3.11Definition of and lower bound of its coarea factor. —
Given an open set , a Lipschitzian map , and , we define by the formula
Clearly, is locally Lipschitzian. If is differentiable at , then so is and, for every , one has
(7) |
Next, we assume that we are given Lipschitzian maps . We define by the formula
It is locally Lipschitzian as well. The relevance of stems from the following observation, assuming that are associated with and as in 3.8 and 3.10:
(8) |
In fact, .
Abbreviate . If each is differentiable at , and , then
where, here and elsewhere, denotes the canonical basis of . Thus, if constitute an orthonormal family in , then
where
(9) |
according to (7), and, in turn,
This allows for a lower bound of the coarea factor of at as follows.
In view of (9), we obtain the next lemma.
3.12. —
Given and , there exists with the following property. Assume that
-
(1)
is open and ;
-
(2)
are Lipschitzian;
-
(3)
is an orthonormal family, for every .
If
-
(4)
, for each ;
-
(5)
;
then
at -almost every .
3.13Definition of and its relation with . —
With , we associate
When is fixed, we also abbreviate as the map defined by . It is then rather useful to observe that, in the context described in 3.8 and 3.10, the following holds:
(10) |
Indeed,
In the sequel, we will sometimes abbreviate . It also helps to notice that, for given and , the set is an -dimensional affine subspace of .
3.14. —
If and , then
is Borel measurable.
Proof.
We start by showing that, when is compact, is upper semicontinuous. Thus, if is a sequence in that converges to , we ought to show that
(11) |
where and . This is indeed equivalent to the same inequality with replaced by , according to 3.2(3) and the last sentence of 3.13. Considering, if necessary, a subsequence of we may assume that none of the compact sets is empty and that the limit superior in (11) is a limit. Since the set of nonempty compact subsets of the compact set , equipped with the Hausdorff metric, is compact, the sequence admits a subsequence (denoted the same way) converging to a compact set . Given , there are converging to . Thus, . In other words, . Therefore, , so that (11) follows from 3.2(1).
Next, we fix and we abbreviate . Thus, we have just shown that contains the collection of all closed subsets of . Observe that if is an nondecreasing sequence in and , then pointwise, thus, . In particular, . Furthermore, if and , then , because all measures involved are finite; indeed, , for all . Accordingly, . This means that is a Dynkin class. Since is a -system, contains the -algebra generated by , i.e., [3, Theorem 1.6.2]. Finally, if , then pointwise, whence is Borel measurable. ∎
3.15. —
Assume and is Borel measurable. The following function is Borel measurable.
Proof.
3.16Definition of . —
Let be Borel measurable and be such that . For each , we define
This is well-defined, according to 3.15. It is easy to check that is a locally finite – hence, -finite – Borel measure on ; indeed, .
To close this section, we discuss the relevance of to the problem of existence of “nearly Nikodým sets”.
3.17Definition of Nearly Nikodým set. —
Let . We say that is nearly -Nikodým in if
-
(1)
;
-
(2)
For -almost each , there is such that .
In case , , and , the existence of such a (with ) was established by Nikodým [12], see also [4, Chapter 8]. For arbitrary and , the existence of such a was established by Falconer [7]. In fact, in both cases, these authors established the stronger condition that, for every , can be replaced by . Thus, in case , letting be a set exhibited by Falconer, if and is such that , then picking arbitrarily such that , we see that . Whence, is also nearly -Nikodým in .
Assuming that is Borel measurable, we say that is nearly -Nikodým in relative to if
-
(1)
;
-
(2)
For -almost each , one has .
3.18. —
Let have finite measure and be Borel measurable. The following are equivalent.
-
(1)
is absolutely continuous with respect to .
-
(2)
There does not exist a nearly -Nikodým set in relative to .
Proof.
A set such that and is, by definition, a nearly -Nikodým set relative to . Condition (1) is equivalent to their nonexistence. ∎
3.19. —
Assume that and that is nearly -Nikodým. The following hold.
-
(1)
There exists Borel measurable such that is nearly - Nikodým in relative to .
-
(2)
There exists compact and continuous such that is nearly -Nikodým in relative to .
Proof.
Define a Borel measurable map by , where is as in 3.7. Choose arbitrarily and define a Borel measurable map
Similarly to (10), observe that
for every . We infer from 3.14 that
is Borel measurable. Thus, the set
is Borel measurable as well. The set is coanalytic and , by assumption. By virtue of von Neumann’s selection theorem [13, 5.5.3], there exists a universally measurable map such that , for every , i.e., . We extend to be an arbitrary constant on . This makes an -measurable map defined on . Therefore, it is equal -almost everywhere to a Borel measurable map . This proves (1).
In order to prove (2), we recall 3.4, specifically, the retraction and the homeomorphic identification . Owing to the compactness of , there are finitely many open balls , , whose closures are contained in and whose union contains . Since , there exists such that , where . It follows from Lusin’s theorem [9, 2.5.3] that there exists a compact set such that and the restriction is continuous. The map takes its values in the closed ball , therefore, it admits a continuous extension . Letting completes the proof. ∎
4. Common setting
4.1Setting for the next three sections. —
In the next three sections, we shall assume the following.
-
(1)
is Borel measurable and .
-
(2)
is open and .
-
(3)
is Borel measurable.
-
(4)
is Lipschitzian.
-
(5)
, for each .
-
(6)
.
-
(7)
and , .
-
(8)
and , .
-
(9)
, for every .
-
(10)
, for every .
-
(11)
constitute an orthonormal basis of , for every .
5. Two fibrations
5.1A fibered space associated with . —
We define
as well as
We will oftentimes abbreviate . It is obvious that is locally Lipschitzian and, therefore, is countably -rectifiable and -measurable. We also consider the two canonical projections
as well as the set
which also is, clearly, countably -rectifiable and -measurable. With the prospect of applying the coarea formula to and , and to and , respectively, we observe that, for each fixed ,
so that
(12) |
and that, for each fixed ,
according to (8), so that
(13) |
whenever . It now follows from the coarea formula that
(14) |
and
(15) |
For these formulæ to be useful, we need to establish bounds for the coarea Jacobian factors and . In order to do so, we notice that if , is differentiable at , i.e., each is differentiable at , , then the approximate tangent space exists and is generated by the following vectors of :
where is an arbitrary basis of . As usual, denotes the canonical basis of .
5.2Coarea Jacobian factor of . —
For -almost every , one has
Proof.
We recall 3.3. That the right hand inequality be valid follows from . Regarding the left hand inequality, fix such that is differentiable at and let denote the restriction of to . Define , , and , . Put
, and
. Recall (6) that
since , . Now, notice that, for ,
whereas, for ,
Since , one also has
Finally,
and the conclusion follows. ∎
5.3. —
Let be an integer and let be an orthonormal family in . There exists such that
Here, denotes the set of increasing maps .
Proof.
We define a linear map and we observe that is an isometry. Therefore, its area Jacobian factor , by definition. Now, also
according to the Binet-Cauchy formula [6, Chapter 3 §2 Theorem 4]. The conclusion easily follows. ∎
5.4Coarea Jacobian factor of . —
The following hold.
-
(1)
For -almost every , one has
-
(2)
For -almost every , one has .
Proof.
Clearly, . Regarding the left hand inequality, fix such that is differentiable at and, this time, let denote the restriction of to . We will now define a family of vectors belonging to . We choose , for . For choosing the remaining vectors, we proceed as follows. We select as in 5.3 applied with to and we let , . Recalling (6), we have
As in the proof of 5.2, we find that
and it remains only to find a lower bound for . The latter equals the absolute value of the determinant of the matrix of coefficients of , , with respect to any orthonormal basis of . We choose the basis . Thus,
(16) |
Abbreviate
and observe that , (recall the proof of 5.2). It remains only to remember that has been selected in order that
and to infer from the multilinearity of the determinant that
This completes the proof of conclusion (1).
Let denote the subset of consisting of those such that each , , is differentiable at . Thus, is Borel measurable and so is
If , then the restriction of to is surjective and, therefore, . Thus, we ought to show that . Since is Lipschitzian, it suffices to establish that . As is Borel measurable, it is enough to prove that , for every , according to Fubini’s theorem. Fix . As in the proof of conclusion (1), choose associated with , according to 5.3. Based on (16), we see that
The set on the right is of the form , for some polynomial , and . It follows that – see, e.g., [9, 2.6.5] – and the proof of (2) is complete. ∎
5.5 Proposition. —
The measure is absolutely continuous with respect to .
Proof.
5.6Definition of . —
Note that is a -finite Borel measure on (see 3.16) and it is absolutely continuous with respect to (see 5.5). It then ensues from the Radon-Nikodým theorem that there exists a Borel measurable function
such that, for every , one has
Furthermore, is univoquely defined (only) up to an null set. This will not affect the reasoning in this paper. Each time we write , we mean one particular Borel measurable function satisfying the above equality, for every .
5.7Definition of . —
5.8 Proposition. —
Given , there exists with the following property. If , then
for -almost every .
Proof.
5.9Rest stop. —
The above upper bound for is already enough to bound it from above, in turn, by a constant times – see 6.4. We next want to establish that , almost everywhere in . Yet, in the definition (17) of , does not appear as the covariable of a function whose level sets we are measuring, thereby preventing the use of the coarea formula in an attempt to estimate . This naturally leads to adding a variable to the fibered space , a covariable for .
5.10A fibered space associated with . —
Let . Abbreviate , the Euclidean ball centered at the origin, of radius , in . We define
and
We note that is locally Lipschitzian and is countably -rectifiable and -measurable. Similarly to 5.1, we define
which, clearly, is also countably -rectifiable and -measurable. We aim to apply the coarea formula to and to the two projections
and
To this end, we notice that
and, thus,
for every . We further notice that
because
and, therefore,
whenever and .
It now follows from the coarea formula and Fubini’s theorem that
(20) |
and
(21) |
5.11Coarea Jacobian factors of and . —
The following inequalities hold, for -almost every .
and
Proof.
The second conclusion is obvious, since . Regarding the first conclusion, we reason similarly to the proof of 5.2. Fix such that is differentiable at and denote by the restriction of to . This tangent space is generated by the following vectors of :
The range of being -dimensional, we need to select vectors in in view of obtaining a lower bound
(22) |
Our choice of is as follows. As in the proof of 5.2, we let , for , and , for . For , we define
for , we define
and, for , we define
so that is an orthonormal basis of and, therefore, the numerator in (22) equals 1. In order to determine an upper bound for its denominator, we start by fixing , we abbreviate and , and we notice that , . Furthermore, since , one has . Therefore, if , then
whereas, if , then
and, if , then
We conclude that
and the proof is complete. ∎
5.12Definition of . —
It follows from the coarea theorem that the function
is -measurable (recall 5.10 applied with ). It now follows from Fubini’s theorem that, for each , the function
is -measurable. In turn, the function
is -measurable. It is a replacement for defined in 5.7. We shall establish, for , a similar lower bound to that in 5.8, this time involving . Before doing so, we notice the rather trivial fact that if , then
for all .
5.13Preparatory remark for the proof of 5.15. —
5.14. —
If is compact, then, for every the function
is upper semicontinuous.
Proof.
The proof is analogous to that of 3.14. For each , define the compact set . If is a sequence converging to , we ought to show that
Since each is a subset of an -dimensional affine subspace of , this is indeed equivalent to the same inequality with replaced by , according to 3.2(3). Considering, if necessary, a subsequence of , we may assume that none of the compact sets is empty and that the above limit superior is a limit. Considering yet a further subsequence, we may now assume that converges in Hausdorff distance to some compact set . One checks that . It then follows from 3.2(1) that . ∎
5.15 Proposition. —
Given , there exists with the following property. If and is compact, then
Proof.
We first observe that we can choose small enough so that
(23) |
for -almost every , provided , according to 5.11. Thus, (23) holds, for -almost every , under the assumption that . In that case, (20), (21), and 5.11 imply that
(24) |
Fix and . According to 5.14, there exists a positive integer such that if , then
Taking the limit superior as , on the right hand side, and letting , we obtain
(25) |
As this holds for all , we may integrate over with respect to . We notice that for every , (recall the notation of 5.13); the latter being -summable, this justifies the application of the reverse Fatou lemma below. Thus, the following ensues from (25), the reverse Fatou lemma, (24), and the Fatou lemma:
∎
5.16 Corollary. —
If and , then
for -almost every .
6. Upper bound for and
6.1Bow tie lemma. —
Let , , and . Assume that
Then there exists such that and . In particular,
Proof.
Let and define . Thus, and, therefore, . Since , we infer that
Therefore, is injective; and the Lipschitzian bound on clearly follows from the above inequality. Regarding the second conclusion, we note that
and is contained in a ball of radius . ∎
6.2. —
Given , there exists with the following property. If
-
(1)
and ;
-
(2)
;
then: For each , each , and each , one has
Proof.
We show that will do. Let , for some . Thus, and, hence,
thus,
In turn,
∎
6.3 Proposition. —
There are and with the following property. If and , then
Proof.
6.4 Corollary. —
There are and with the following property. If , then
for -almost every .
Proof.
Let . ∎
7. Lower bound for and
7.1Setting for this section. —
7.2Polyballs. —
Given and , we define
We notice that, if , then ; in particular, . We also notice that .
7.3. —
Given , there exists with the following property. If
-
(1)
;
-
(2)
;
-
(3)
;
-
(4)
is closed;
then
7.4 Remark. —
With hopes that the following will help the reader form a geometrical imagery: Under the circumstances 7.3, may be seen as a “nonlinear stripe”, “horizontal” with respect to , “at height” with respect to , and of “width” .
Proof of 7.3.
Given , we define
and we consider the isometric parametrization defined by the formula
We also abbreviate .
Claim #1. .
Since is an isometry, it suffices to obtain an upper bound for . Let and note that
Recalling hypothesis (1), it is now apparent that can be chosen small enough, according to , , and , so that Claim #1 holds.
Claim #2. For -almost every , one has .
Let be such that is differentiable at . We shall estimate the coefficients of the matrix representing with respect to the canonical basis. Fix and recall (7):
Next, note that
where the last inequality follows from hypothesis (1), upon choosing small enough, according to , , and . Moreover,
Therefore, if is the matrix representing with respect to the canonical basis, we have shown that , for all . This completes the proof of Claim #2.
Claim #3. .
We shall show that , for every , and the conclusion will become a consequence of the intermediate value theorem, in case , or a standard application of homology theory, see, e.g., [5, 4.6.1], in case . If , it is clearly enough to establish this inequality only for -almost every . In that case, owing to the coarea theorem [9, 3.2.22], we may choose such a so that that is differentiable -almost everywhere on the line segment . Whether , or , it then follows from Claim #2 that
Accordingly,
and the claim will be established upon showing that . Note that ; we shall use hypothesis (3) to bound its norm from above. Given , recall that , thus,
where the last inequality holds, according to hypothesis (1), provided is chosen sufficiently small. In turn,
by virtue of hypothesis (3).
Claim #4. For every and every closed , one has .
First, notice that
and, therefore,
since is an isometry. Now, since , according to Claim #3, we have
It therefore follows from Claim #1 that
We are now ready to finish the proof, by an application of Fubini’s theorem:
∎
7.5Lower bound for . —
Given , there exists with the following property. If
-
(1)
;
-
(2)
;
-
(3)
is closed;
-
(4)
;
then
where is a positive integer depending only upon .
Proof.
Similarly to the proof of 7.3, we will first establish a lower bound for on “vertical slices” of the given polyball, followed, next, by an application of Fubini. Given , we let and be as in 7.3, and we consider the set
(notice is slightly smaller than used in the proof of 7.3) and the isometric parametrization defined by
For part of the proof, we find it convenient to abbreviate . We also put .
By definition of , for each , there exists a collection of closed balls in with the following properties: For every , is a ball centered at 0, ,
and . Furthermore, being -summable, according to 6.3, there exists such that and every is a Lebesgue point of . For such a , we may reduce , if necessary, keeping all the previously stated properties valid, while enforcing also that
whenever . We infer that, for each and each ,
(26) |
It ensues from the Vitali covering theorem that there is a sequence in , and , such that the balls , are pairwise disjoint and . It therefore follows, from (26) and the fact that is an isometry, that
(27) |
where we have abbreviated . We also abbreviate and we infer from the coarea formula that, for each ,
(28) |
where the last inequality follows from 3.12, applied with , provided that is chosen smaller than . Letting and recalling that , we infer, from (27) and (28), that
(29) |
Applying 7.3 to each does not immediately yield a lower bound for , because the are not necessarily pairwise disjoint. This is why we now introduce slightly smaller versions of these:
Claim. The sets , , are pairwise disjoint.
Assume, if possible, that there are and . Letting and denote, respectively, the radius of , and of , we notice that , because . Since is an isometry, we have and, therefore, also
(30) |
We now introduce the following vectors of :
and we notice that
where the second equality holds because , as clearly follows from the definition of . Furthermore,
since we may choose to be so small that the last inequality holds, in view of hypothesis (1). Whence,
contradicting (30). The Claim is established.
7.6 Proposition. —
Given , there exist and with the following property. If
-
(1)
;
-
(2)
;
-
(3)
is closed;
-
(4)
;
then
7.7 Remark. —
The difference with 7.5 is the domain of integration (being smaller) in the integral, on the left hand side in the conclusion.
Proof of 7.6.
The reader will happily check that
suits their needs. ∎
7.8 Proposition. —
There exists with the following property. If , then
for -almost every .
Proof.
We let
According to 5.16, it suffices to show that , for -almost every . Define and assume, if possible, that . Since is -measurable (recall 5.12), there exists a compact set such that . Observe that the sets , for and , form a density basis for -measurable subsets of – because their eccentricity is bounded away from zero – thus, there exists and such that
whenever . There is no restriction to assume that is small enough for . Thus, if we let , it follows from 7.6 that
(33) |
On the other hand, recalling 5.12 and the fact that , we infer that , for all . In particular, , for all , contradicting (33). ∎
8. Proof of the theorems
8.1 Theorem. —
Assume that , is Lipschitzian, and is Borel measurable. The following are equivalent.
-
(1)
.
-
(2)
For almost every , .
-
(3)
For almost every , .
Recall our convention that .
Proof.
Since is complete, we can extend to the closure of . Furthermore, if the theorem holds for , then it also holds for . Thus, there is no restriction to assume that is closed.
. It follows from 3.8 that each admits an open neighborhood in such that can be associated with a Lipschitzian orthonormal frame satisfying all the conditions of 4.1, for some . Since is Lindelöf, there are countably many such that . Letting , we infer from 5.5 that is absolutely continuous with respect to . Thus, if , then , for -almost every , by definition of . Since is arbitrary, the proof is complete.
is trivial.
. Let satisfy condition (2). It is enough to show that , for each . Fix and define . Consider the defined in the second paragraph of the present proof; since is compact, finitely many of those, say , cover . Let . Partition each , , into Borel measurable sets , , such that . It then follows from 7.8 that
(34) |
for -almost every . Now, fix and . Observe that , for -almost every . Thus, . Moreover,
It follows from (34) that . Since and are arbitrary, . ∎
8.2 Remark. —
Alternatively, one can prove the principal implication in two other ways. One way – more involved – consists in applying our main result 8.4 below. A second – simpler – way, along the following lines, avoids reference to the estimates we obtained for the functions and . Consider , , and such that , and put . Define
is locally Lipschitzian and one checks that, in fact, is a lipeomorphism between and , where , for some depending upon and , because its differential is close to the identity. Referring to Fubini’s theorem, one then further checks that
Finally, using Fubini again, with respect to the decomposition , one shows that , for -almost every , where is as close as we wish to . Applying the previous construction with replacing , using , we find that , for some depending on .
Notwithstanding, it seems that the (simpler) change of variable described here is not enough to yield the (stronger) theorem below.
8.3Polyballs. —
Recalling 7.2, we notice that
where is a norm on defined by the formula
for . It is readily observed that .
-
(1)
One has , for -almost every .
Abbreviate and and define , so that . Let and notice is differentiable at . We henceforth assume that , whence, – the proof in the other case is similar. Define , , and let . Note that and . Thus, and, in turn, . It follows that . Since is a unit vector and , we conclude that .
-
(2)
Let , and be as in 4.1. Assume , , and is so that . It follows that
First notice that , for every , so that and, for every , . Next notice that, for each ,
We now assume that , in particular, , whence,
8.4 Theorem. —
Assume that is Borel measurable and that is Lipschitzian. It follows that
for -almost every .
Recall our convention that .
Proof.
Extend to in a Borel measurable way, for instance, to be an arbitrary constant outside of . It follows from 3.15 that, for each , the function is Borel measurable. Thus, for each , the function defined by
is Borel measurable as well, and so is .
Abbreviate . Arguing reductio ad absurdum, we henceforth assume that and fail the conclusion of the theorem. Thus, the set is Borel measurable and non Lebesgue null. Accordingly, there exists such that the set is also Borel measurable and of positive Lebesgue measure. It therefore ensues from Egoroff’s theorem [9, 2.3.7] that there exists a closed set such that and that there exists a positive integer such that
(35) |
for each and each . Choose such that
(36) |
and choose such that
(37) |
As in the proof of 7.8, we recall that the family , for and , is a density basis of -measurable sets. Since , there exists such that
In particular, there exists such that
(38) |
whenever .
References
- [1] J. Bourgain, A remark on the maximal function associated to an analytic vector field, Analysis at Urbana, Vol. I (Urbana, IL, 1986–1987), London Math. Soc. Lecture Note Ser., vol. 137, Cambridge Univ. Press, Cambridge, 1989, pp. 111–132. MR 1009171
- [2] A. M. Bruckner, Differentiation of integrals, Amer. Math. Monthly 78 (1971), no. 9, part II, ii+51. MR 293044
- [3] D. L. Cohn, Measure theory, second ed., Birkhäuser Advanced Texts: Basler Lehrbücher. [Birkhäuser Advanced Texts: Basel Textbooks], Birkhäuser/Springer, New York, 2013. MR 3098996
- [4] M. de Guzmán, Real variable methods in Fourier analysis, North-Holland Mathematics Studies, vol. 46, North-Holland Publishing Co., Amsterdam-New York, 1981, Notas de Matemática [Mathematical Notes], 75. MR 596037
- [5] Th. De Pauw, Concentrated, nearly monotonic, epiperimetric measures in Euclidean space, J. Differential Geom. 77 (2007), no. 1, 77–134. MR 2344355
- [6] L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions, Studies in Advanced Mathematics, CRC Press, Boca Raton, FL, 1992. MR 1158660 (93f:28001)
- [7] K. J. Falconer, Sets with prescribed projections and Nikodým sets, Proc. London Math. Soc. (3) 53 (1986), no. 1, 48–64. MR 842156
- [8] H. Federer, Curvature measures, Trans. Amer. Math. Soc. 93 (1959), 418–491. MR 110078
- [9] by same author, Geometric measure theory, Die Grundlehren der mathematischen Wissenschaften, Band 153, Springer-Verlag New York Inc., New York, 1969. MR 0257325 (41 #1976)
- [10] A. B. Kharazishvili, Nonmeasurable sets and functions, North-Holland Mathematics Studies, vol. 195, Elsevier Science B.V., Amsterdam, 2004. MR 2067444 (2005d:28001)
- [11] M. Lacey and X. Li, On a conjecture of E. M. Stein on the Hilbert transform on vector fields, Mem. Amer. Math. Soc. 205 (2010), no. 965, viii+72. MR 2654385
- [12] O. Nikodým, Sur la mesure des ensembles plans dont tous les points sont rectilinéairement accessibles., Fundam. Math. 10 (1927), 116–168 (French).
- [13] S. M. Srivastava, A course on Borel sets, Graduate Texts in Mathematics, vol. 180, Springer-Verlag, New York, 1998. MR 1619545 (99d:04002)