On the heterogeneous distortion inequality
Abstract.
We study Sobolev mappings , , that satisfy the heterogeneous distortion inequality
for almost every . Here is a constant and is a function in . Although we recover the class of -quasiregular mappings when , the theory of arbitrary solutions is significantly more complicated, partly due to the unavailability of a robust degree theory for non-quasiregular solutions. Nonetheless, we obtain a Liouville-type theorem and the sharp Hölder continuity estimate for all solutions, provided that for some . This gives an affirmative answer to a question of Astala, Iwaniec and Martin.
Key words and phrases:
Heterogeneous distortion inequality, Quasiregular mappings, Liouville theorem, Hölder continuity, Astala-Iwaniec-Martin question2020 Mathematics Subject Classification:
Primary 30C65; Secondary 35B53, 35R45, 53C211. Introduction
Let be a connected, open subset of with . We study mappings in the Sobolev space that satisfy the heterogeneous distortion inequality
(1.1) |
for a.e. (almost every) , where and . Here, is the operator norm of the weak derivative of at a point ; that is, . Moreover, is the Jacobian determinant of . It is worth noting that the natural Sobolev space in which to seek the solutions is . The reason for this is that this space provides enough regularity to apply integration by parts to the form .
If in (1.1), we recover the mappings of bounded distortion, also known as -quasiregular maps. Homeomorphic -quasiregular maps are also commonly called -quasiconformal. The theory of mappings of bounded distortion arose from the need to generalize the geometry of holomorphic functions to higher dimensions, and is by now a central topic in modern analysis with important connections to partial differential equations, complex dynamics, differential geometry and the calculus of variations; see the monographs by Reshetnyak [32], Rickman [33], Iwaniec and Martin [19], and Astala, Iwaniec and Martin [1]. The last 20 years have also seen widespread study of a more general class of deformations, mappings of finite distortion, where the constant is replaced by a finite function ; see e.g. the monographs [19] and [14].
A core part of the theory of quasiregular mappings is that the distortion estimate implies several strong topological properties. In higher dimensions this was pioneered in a series of papers by Reshetnyak (1966–1969). Accordingly, spatial quasiregular mappings enjoy the following properties:
-
(i)
A -quasiregular mapping is locally -Hölder continuous [28];
- (ii)
-
(iii)
A bounded quasiregular mapping in is constant [29].
Recall that a mapping is open if is open for every open , and that is discrete if the sets consist of isolated points. Generalizations of these results also hold for mappings of finite distortion, assuming sufficient integrability conditions on the distortion function; see e.g. [35, 17, 21, 34, 15].
Solutions to the heterogeneous distortion inequality (1.1) generally lack these powerful conditions, and therefore require new tools for their treatment. For instance, if in and , then satisfies (1.1) with . This results in a discontinuous solution by simply choosing a discontinuous . Similarly, by choosing a smooth, compactly supported , we have that satisfies (1.1) with for every , yet is neither constant nor discrete or open. Hence, the heterogeneous distortion inequality even with a regular is too weak for its solutions to satisfy the above conditions (ii) and (iii), and therefore the use of e.g. degree theory [9] is unavailable for non-quasiregular solutions.
However, in the planar case , Astala, Iwaniec and Martin showed that entire solutions of (1.1) which vanish at infinity do satisfy a counterpart of the Liouville theorem (iii), provided that is sufficiently regular; see [1, Theorem 8.5.1]. This uniqueness theorem has found several important applications, such as the nonlinear -problem in the theory of holomorphic motions [22] or the solution to the Calderón problem [2]. It is for this reason that Astala, Iwaniec and Martin asked in [1, p. 253] whether their form of the Liouville theorem and a version of the continuity result (i) remain valid in higher dimensions.
Astala-Iwaniec-Martin Question. Suppose that satisfies the heterogeneous distortion inequality (1.1) with for some . Under these assumptions, is the mapping continuous? Moreover, does satisfy the following Liouville-type theorem: if as , then ?
Along with their proof of the planar case, Astala, Iwaniec and Martin provided a counterexample which shows that the assumption is not enough for the Liouville-type theorem; see [1, Theorem 8.5.2.]. Note that in the planar case, the heterogeneous distortion inequality (1.1) with amounts to saying that the solutions satisfy the homogeneous differential inequality
(1.2) |
with and . Moreover, (1.2) can in fact be expressed as a linear heterogeneous Cauchy-Riemann system, and is thus uniformly elliptic. However, for , the theory is nonlinear, and the main inherent difficulty lies in the lack of general existence theorems and counterparts to power series expansions of the solutions.
In this paper, we give an affirmative answer to the Astala–Iwaniec–Martin question. The continuity is the more straightforward part of the solution, whereas the main challenge lies in proving the Liouville-type uniqueness theorem.
1.1. Continuity
The solutions to (1.1) are indeed continuous when . We in fact establish sharp local Hölder continuity estimates for such mappings.
For we denote the class of locally -Hölder continuous mappings by . We let denote the sharp Hölder exponent of -quasiregular mappings, and the sharp Hölder exponent of -functions. Our result shows that if , then the sharp Hölder exponent for solutions to (1.1) is the minimum of the exponents and . There is, however, a somewhat surprising special case: if and coincide, then we in fact end up with an infinitesimally weaker Hölder exponent than .
Theorem 1.1.
Let be a domain, , , , and . Suppose that satisfies the heterogeneous distortion inequality (1.1) with and .
If , then . Moreover, there exist and satisfying the assumptions such that for any .
If , then for every , and there exist and satisfying the assumptions such that .
1.2. Liouville Theorem
Recall that the classical Liouville theorem asserts that bounded entire holomorphic functions are constant [5]. Its quasiregular counterpart (iii) follows from the Caccioppoli inequality [3], which controls the derivatives of a quasiregular mapping locally in terms of the mapping. Attempting the same approach under the assumptions of the Astala-Iwaniec-Martin question yields that the integral of the Jacobian over the entire space equals zero, or equivalently, that
(1.3) |
Analogously to the quasiregular theory [10], the local variants of the energy estimate (1.3) imply a higher degree of integrability for the derivatives of a solution. These observations can be further combined with the nonlinear Hodge theory developed by Iwaniec and Martin [18, 16], which provides -estimates for with exponents either smaller or larger than the dimension . However, this approach alone appears insufficient for the desired Liouville-type theorem, and although it could be used to show the continuity of , the proof would not yield the sharp Hölder exponents we obtain in Theorem 1.1.
Our first Liouville-type result shows that if , then nontrivial solutions to (1.1) that are bounded in do not vanish at any point. Note that the standard radial example of a -quasiregular mapping shows that this result has no local alternative.
Theorem 1.2.
Suppose that satisfies the heterogeneous distortion inequality (1.1) with and for some . If is bounded and is not identically zero, then .
The first and key step in the proof of Theorem 1.2 is to show that a solution to the heterogeneous distortion inequality with satisfies
(1.4) |
The proof of this is based on a thorough analysis of the integrals of over the sublevel sets of . In turn, we obtain that . The second step is to establish a Morrey-type decay estimate [24] on the integrals of over the balls in . Our proof of the decay estimate relies on the formal identity
where is a certain differential -form. The resulting polynomial decay estimate implies that the function is Hölder continuous. Thus, the mapping itself must omit the point from its range, completing the proof of Theorem 1.2.
Our final main result is then the Liouville part of the Astala-Iwaniec-Martin question.
Theorem 1.3.
Suppose that satisfies the heterogeneous distortion inequality (1.1) for and for some . If , then .
We prove Theorem 1.3 by showing that, if there exists a non-identically zero solution to (1.1) with , then the oscillation of the function over the entire space is uniformly bounded. This clearly leads to a contradiction with the assumption when . The crucial step in obtaining the uniform oscillation bound is to strengthen the estimate (1.4); that is, to prove an integrability estimate below the natural exponent for the expression .
Our solution is influenced by the case . Indeed, in this case, the mapping has no zeros by Theorem 1.2, and hence has a well-defined complex logarithm . The mapping satisfies the distortion estimate
(1.5) |
almost everywhere. This, in turn, gives a nonhomogeneous linear elliptic equation for , which implies the desired integrability estimate below the natural exponent for . In higher dimensions, the issue is to construct a similar map . The Zorich map provides a well-known -dimensional generalization of the planar exponential mapping; see [37]. Unfortunately, has a branch set consisting of -dimensional hyperplanes, which prevents lifting an arbitrary continuous through .
We circumvent the lifting difficulties by moving to the Riemannian manifold setting. Indeed, a well defined counterpart for the logarithm exists from to , providing us with a mapping . This mapping satisfies a higher dimensional counterpart of the estimate (1.5), and the single Euclidean component in the target space is sufficient for a Caccioppoli-type inequality to hold, which leads to the desired integrability estimate below the natural exponent .
Acknowledgments
We thank Tadeusz Iwaniec and Xiao Zhong for discussions and shared insights.
2. Preliminaries
In this section, we go over some of the tools we require which might be less familiar to readers.
2.1. Sobolev mappings with manifold target
For the most part of this text we use the standard Sobolev spaces and , where is an open connected set, see e.g. [9, 36]. However, towards the end of the text, we consider a locally Sobolev mapping , where is an -dimensional Riemannian manifold.
There are various approaches to defining first order Sobolev mappings with a manifold target; see e.g. [11] or [6]. However, in our case, we only have to consider continuous Sobolev mappings with a manifold target, which simplifies the definition significantly.
Definition 2.1.
Let be a domain, and let be a Riemannian -manifold. We say that a continuous is in the Sobolev space for if, for every , there exists a neighborhood of and a smooth bilipschitz chart such that and .
If a continuous function is in , then there exists a weak derivative which satisfies for bilipschitz charts . This weak derivative is unique up to a set of measure zero, in the sense that if is another such mapping, then outside a set of the form where the set has zero -dimensional Lebegue measure, .
At a given point , we denote by the operator norm of , where is equipped with the norm induced by the Riemannian metric. It follows that is in for any continuous . If and is oriented, then we also have a measurable Jacobian , characterized almost everywhere by .
We remark that if , then the above definition coincides with the usual definition of for continuous . We also remark that if the target is a product manifold , then given two continuous mappings and , we have that if and only if and .
Next we recall Sobolev differential forms. Namely, suppose that is a Riemannian manifold. A measurable -form is the weak differential of a measurable -form if
for every . We denote by the space of -forms with a weak differential . The version where and is denoted . We also use the shorthands and .
In particular, we require the following standard result about pull-backs of compactly supported smooth forms with Sobolev mappings. We sketch the proof for the convenience of the reader.
Lemma 2.2.
Let be a domain, and let be a Riemannian -manifold. Suppose that , where we assume that is continuous if . If and , then and .
Sketch of proof.
The fact that and satisfy the correct integrabilities follows from the estimates
for a.e. .
For , it suffices to consider for which the support of is contained in the domain of a bilipschitz chart , as the general is a finite sum of such forms. Moreover, it suffices to consider of the form , as a general with is again a finite sum of such forms with the coordinates of rearranged. We may also select such that on , which lets us write in a form where the components are defined on all of .
We then use the chain rule of locally Sobolev and maps to conclude that, , and for every . By the wedge product rules for Sobolev forms and the formula for the weak differential, we conclude that . ∎
2.2. Caccioppoli inequality
The Caccioppoli inequalities are a standard tool in the study of quasiregular mappings. The most basic form of the Caccioppoli estimate for a -quasiregular mapping reads as
where is a real-valued smooth test function with compact support in . This follows from the general inequality
which can be proved for arbitrary via an integration-by-parts argument. Our arguments, however, require a version with a target space other than . In general, it is not possible to obtain a Caccioppoli-type estimate for mappings where is a Riemannian -manifold. This happens when is a rational homology sphere, see [11]. However, the standard proof generalizes to the case of , where is a Riemannian -manifold.
Lemma 2.3.
Let be a domain, and let be a compact oriented Riemannian -manifold without boundary. Let , where we assume that is continuous if . Denote by and the coordinate functions of . Then for every and every , we have
Proof.
We define the function . Then . Moreover, since , we have
We may select a subdomain with smooth boundary such that is compactly contained in and . Since with compact support, we may approximate it in with . Moreover, we have by Lemma 2.2 that and . Hence, , and we may approximate in with (see e.g. [20, Corollary 3.6]).
By a standard Hölder-type estimate and the Leibniz rule, it therefore follows that in . However, since is smooth and compactly supported for every , it follows that
Hence, we may estimate that
The claim therefore follows. ∎
2.3. Jacobians of entire mappings
To end the preliminaries section, we also discuss a result regarding the Jacobian of an entire Sobolev mapping. It is the main reason why we stated the Caccioppoli inequality also in the case where the target is the product of and an -manifold.
Lemma 2.4.
Let be a compact oriented Riemannian -manifold without boundary. Suppose that , where we assume that is continuous if . If , then
Proof.
We let be such that we have , , and . We again denote . We then use the Caccioppoli inequality of Lemma 2.3 and Hölder’s inequality to obtain
Since and , the Sobolev-Poincaré inequality then yields that
Since , the first integral term on the right hand side tends to 0 as , while the second term stays bounded. Since , the claim therefore follows by dominated convergence. ∎
3. Hölder continuity
In this section, we prove the continuity part of Theorem 1.1. Our proof is based on Morrey’s rather elegant ideas in geometric function theory [24, 28, 19]. A crucial tool in establishing the sharp Hölder exponent is the isoperimetric inequality in the Sobolev space . For and almost every such that compactly contained in , we have
(3.1) |
where is the -dimensional area of the unit sphere in . Here stands for the cofactor matrix of the differential matrix . For a diffeomorphism this integral form of the isoperimetric inequality follows immediately from the familiar geometric form of the isoperimetric inequality
where stands for the volume of a domain and is its -dimensional surface area. For the proof of (3.1) for Sobolev mappings see Reshetnyak [32, Lemma II.1.2.] for a more detailed account.
We begin with the primary estimate our proof relies on.
Lemma 3.1.
Let be bounded domains with . Suppose that satisfies the heterogeneous distortion inequality (1.1) for and , where . Let , let for all , let be such that . Then for every and a.e. we have the estimate
where depends only on , , , , and . In particular, doesn’t depend on , and . Moreover, if , then the estimate also holds for .
Proof.
By using the heterogeneous distortion inequality, the isoperimetric inequality (3.1) for -mappings, Hadamard’s inequality , and Hölder’s inequality, we obtain for a.e. the estimate
For the final term, we note that for every , we have by the Sobolev embedding theorem, and consequently also . Hence, we may use Hölder’s inequality to obtain the desired estimate
Moreover, if we additionally know that , we obtain the claim for by estimating
∎
Note that the estimate of Lemma 3.1 is of the form . This differential inequality allows us to obtain an estimate for the decay of at 0, which in our case is a decay estimate on the integrals of over balls.
Lemma 3.2.
Suppose that is an absolutely continuous increasing function such that and
(3.2) |
for a.e. , where and . Then there exists a constant such that the following holds:
-
•
if , then for all we have
-
•
if , then for all we have
-
•
if , then for all we have
Proof.
We observe that
for a.e. . Consequently, the estimate (3.2) can be rewritten in the form
We integrate this estimate, obtaining
(3.3) |
Consider first the case . Computing the integrals in (3.3) yields
and further rearrangement and estimation yields
Suppose then that . Then (3.3) results in
and we may again further estimate
Finally, consider the case . In this case, it follows from (3.3) that
and further rearrangement and estimation yields
∎
For the remaining component to the proof of Theorem 1.1, we recall a well known fact that the decay estimate on implies that belongs to a Morrey–Campanato space [25, 4], and is thus Hölder continuous. The precise formulation of this fact that we use is as follows.
Lemma 3.3.
Let for some and . Suppose that satisfies
(3.4) |
for all , where , , and is large enough that . Then
for all , where depends on , , , , and .
Note that the assumption is to ensure that is increasing on . Lemma 3.3 is merely a small variant of a classical result of Morrey [24] with an extra logarithmic term, where the logarithmic term becomes relevant when investigating the exact modulus of continuity. See [26, Theorem 3.5.2] for a proof in the classical case . For general , we note that Lemma 3.3 also follows from the fractional maximal function estimate of Sobolev functions: if and , then for all outside a set of measure zero, we have
(3.5) |
where stands for the restricted fractional maximal function
Indeed, taking close to and combining (3.4) with (3.5) yields the desired estimate of Lemma 3.3. The proof of (3.5) is due to Hedberg [12].
We are now ready to prove the local Hölder continuity stated in Theorem 1.1.
Lemma 3.4.
Let for some . Suppose that satisfies the heterogeneous distortion inequality (1.1) with and , where .
If , then
for all , where .
If , then
for all , where .
Proof.
We first prove a slightly weaker Hölder continuity estimate for than is claimed. This in turn implies the local boundedness of , which lets us apply Lemma 3.1 in full force and to obtain the stated estimates.
We set and choose . Applying Lemmas 3.1 and 3.2 we conclude that
(3.6) |
for all . Therefore, it follows from Lemma 3.3 that is -Hölder continuous in . Since continuity is a local property, we conclude that is continuous, and in particular bounded in .
Now, knowing that is locally bounded we may and do take in Lemma 3.1. Combining this with Lemma 3.2, we obtain the following decay estimate for the differential:
(3.7) |
for all . Thus, the desired Hölder continuity estimates for follow from Lemma 3.3.
∎
4. Sharpness of the Hölder exponents
Having Lemma 3.4, the remaining part of proving Theorem 1.1 is to construct solutions which show that the obtained Hölder exponents cannot be improved. Recalling the notation and , we consider three different cases: , and .
For the first case , we can simply use the standard radial example
Indeed, the mapping is -quasiregular and hence satisfies (1.1) with , and we also have for every .
Next, we discuss the case in-depth.
Example 4.1.
Let and such that . We define a mapping with only a single non-vanishing coordinate function, namely
This mapping lies in , with
Furthermore, . Hence, the heterogeneous distortion inequality (1.1) for reduces to
Since for every , the mapping solves the heterogeneous distortion inequality for any . We choose
and then observe that
We recall that for any , the function is integrable over . Indeed,
Hence, the mapping solves (1.1) with . However, for any exponent , the map fails to be -Hölder continuous at the origin.
The remaining part of the proof of Theorem 1.1 is therefore to provide an example in the special case .
Example 4.2.
Let and , and suppose that . We define a mapping by
The mapping is hence obtained by shifting a radially symmetric map of the form , where . For we have
Using these and the fact that is orientation preserving, we conclude, see e.g. [19, 6.5.1], that
and
Since is increasing on , we have for all . Therefore, the heterogeneous distortion inequality (1.1) is satisfied if we choose
We then observe that
and since
we conclude that . However, we have
and therefore .
The proof of Theorem 1.1 is thus complete.
5. Sublevel sets and the logarithm
In this section, we begin studying bounded entire functions satisfying (1.1), with the goal of eventually reaching the Liouville type theorem stated in the Astala-Iwaniec-Martin question. Our main goal in this section is to show that if is not identically zero, then . This is already notable, since this condition is not satisfied by all unbounded entire quasiregular maps. Our approach does not rely on the theory of partial differential equations. Instead, the proof is based on two main tools: integration by parts and truncating with respect to its level sets.
5.1. Global integrability
We begin with a simple global integrability result for when is an entire mapping that solves the heterogeneous distortion inequality (1.1).
Lemma 5.1.
Suppose that satisfies the heterogeneous distortion inequality (1.1) with and . If is bounded, then .
Proof.
Let be a smooth mapping chosen such that , , and . Now, by using the heterogeneous distortion inequality, the Caccioppoli estimate of Lemma 2.3, and Hölder’s inequality, we obtain
Hence, we obtain an upper bound on the integral of independent on . Letting yields the claim. ∎
5.2. Level set methods
We just proved that for a bounded satisfying (1.1) with , the differential lies in . Therefore, by Lemma 2.4, the integral of over the entire space is zero. We now proceed to improve this by showing that the integral of the Jacobian also vanishes over every strict sublevel set of .
Lemma 5.2.
Let . Suppose that . Then for every , we have
Proof.
Let , and let be a non-decreasing smooth function such that , , and . Let be the radial function defined by
Then is a smooth and 2-Lipschitz regular mapping. Consequently, the chain rule applies, for a.e. , and lies in , see e.g. [7, p.130].
In particular, since , we have . Therefore, Lemma 2.4 yields that
As , we have pointwise where denotes the characteristic function of a set . Hence, the claim follows by letting and applying the dominated convergence theorem for the Lebesgue integral. ∎
Lemma 5.2 is our main tool in showing that, for an entire non-identically zero solution , the function belongs to . Towards this, we first prove that the function is globally integrable.
Lemma 5.3.
Suppose that solves the heterogeneous distortion inequality (1.1) with and . If is bounded, then is integrable,
and |
Here and in what follows, we interpret when , and similarly, when .
Proof.
We split into its positive and negative parts . These parts satisfy the inequality
(5.1) |
In particular, we have
(5.2) |
Indeed, this is trivial when , and if , then (5.2) follows from (5.1). Hence, a.e. where , and we have
(5.3) |
For every , we denote the strict sublevel set of at by . By Lemmas 5.1 and 5.2, we have
In particular, we have for every that
Multiplying this estimate by , we obtain
(5.4) |
By integrating (5.4) over with respect to and applying the Fubini–Tonelli theorem to change the order of integration, we have
Evaluating the inner integral yields
where the finiteness of the integrals follows from (5.3).
We therefore conclude that a.e. where , that and therefore also are integrable, and that
Finally, by the heterogeneous distortion inequality (1.1), we see that a.e. where , and that
∎
5.3. The logarithm
With the global integrability of shown, we now proceed to study the Sobolev regularity of .
Let be a ball in with . Suppose that , and that . For every , we denote by the function . We then proceed to study the functions . Since is bounded and the function given by is locally Lipschitz, we may use the chain rule of Lipschitz and Sobolev maps to obtain that ; see e.g. [36, Theorem 2.1.11]. Moreover, we have the uniform estimate
(5.5) |
which is independent of .
By using these truncated logarithms as a tool, we achieve the following result.
Lemma 5.4.
Let be a bounded and not identically zero mapping. Suppose that and that Then has zero Lebesgue measure, the measurable function lies in , and
Proof.
By our assumptions, the set has positive measure. Hence, there exists such that has positive measure. For every , we denote by the function .
Our first goal is to show that . For the proof, we assume towards a contradiction that the integral of over is instead infinite. In this case, since the functions are uniformly bounded from above and decrease to monotonically as , we have ; recall that stands for the integral average value of the function over .
By the Sobolev-Poincaré inequality and (5.5), we have the upper bound
This upper bound, independent of , is finite by our assumptions. We also have the lower bound
Since , we arrive at a contradiction. Hence, . In particular, it follows that is finite almost everywhere, and therefore has zero Lebesgue measure.
Now, for , we have
and
Therefore, in and in . Thus, the weak gradient of equals . Since , the Sobolev embedding theorem shows that , and hence . ∎
6. Non-existence of zeroes
In this section we will show that is locally Hölder continuous if is a bounded entire solution to the heterogeneous distortion inequality with . This will prove Theorem 1.2. Our approach again mimics the lines of reasoning by Morrey and is based on obtaining a quantitative integral estimate for over balls. This is done by employing a suitable isoperimetric inequality.
6.1. Logarithmic isoperimetric inequality
We then proceed to show the following isoperimetric-type estimate for . As before, we use to denote a ball in around a fixed point .
Lemma 6.1.
Let , where . If , then there is a constant such that for a.e. , we have
(6.1) |
The main idea behind the proof is to write
where is a certain differential -form, and then to use Stokes’ theorem. This method actually gives similar estimates for integrals of the more general form over balls. The precise estimate obtained is given by the following lemma, which is a variant of [27, Lemma 2.1] by Onninen and Zhong. We provide a proof here due to our assumptions being slightly weaker than in [27].
Lemma 6.2.
Let be a domain. Suppose that is in . If is a piecewise -smooth function with locally bounded, then for every test function , we have
Before jumping into the proof, we comment on the well-definedness of the left integrand. The function is defined outside finitely many jump points . Consider the set . Then a.e. on , we have ; see e.g. [13, Corollary 1.21]. Since , this implies that for a.e. , we have or some non-zero linear combination of vanishes. In the latter case, . Hence, almost everywhere where is not defined, we have or , making the integrand in the statement well defined.
Proof of Lemma 6.2.
By switching to a smaller domain which still contains the support of , we may assume that is bounded and . By boundedness of and the local boundedness of and , we may also assume that and are bounded.
We consider the function
Since is bounded and is Sobolev, the chain rule of Lipschitz and Sobolev maps yields that a.e. on , see e.g. [36, Theorem 2.1.11]. By further using the product rules of Sobolev mappings, we see that has a locally integrable weak gradient given by
Using the fact that , , and are bounded, we then conclude that this weak gradient is in .
We therefore have that . Consequently, , and therefore is integrable. Since also has a compactly supported coordinate function, we therefore have
By writing as a wedge product, we obtain
By summing over , and by using the fact that for 1-forms , the claim follows. ∎
Proof of Lemma 6.1.
We first prove an isoperimetric estimate of the following form: for a.e. and every constant , we have
(6.2) |
Hence, fix a and let . For all sufficiently large , we select cutoff functions such that , for all , and . We also fix and , and define a function by
The function is piecewise and its derivative is locally bounded. Moreover, we have
Hence, by using Lemma 6.2 with and , we obtain that
where we recall that . We then let , where we use monotone convergence and the Lebesgue differentiation theorem to obtain
for a.e. . Combining this with Hadamard’s inequality yields
(6.3) |
for a.e. .
Next, we let . Since is integrable and has zero measure by Lemma 5.4, the last integral in (6.3) goes to zero as . For the first integral on the right hand side of 6.3, we observe that its integrand is dominated by the function for some . This dominant is in for any , since , and since by Lemma 5.4. Consequently, the dominant is also in for a.e. by the Fubini-Tonelli theorem. Hence, we may apply the dominated convergence theorem as in (6.3), and therefore obtain
for a.e. .
We then let and again use the dominated convergence theorem, obtaining the claimed inequality (6.2) for a.e. and our fixed value of . Consequently, if is a countable dense subset, then (6.2) holds for a.e. and all . We then obtain (6.2) for all and a.e. by taking limits, since the constant in (6.2) is independent of and since is integrable over for a.e. .
It remains to derive the statement of the lemma from (6.2). We denote
Since , the Sobolev embedding theorem on spheres [14, Lemma 2.19] implies that, after changing in a set of measure zero, we have
(6.4) |
for a.e. . Moreover, if is such that (6.2) is valid, we may select and take , in which case (6.2) yields
(6.5) |
for a.e. .
6.2. Hölder continuity of the logarithmic function
Here we complete the proof of Theorem 1.2. We recall the statement first.
Theorem 1.2.
Suppose that satisfies the heterogeneous distortion inequality (1.1) with and where . If is bounded and , then .
The proof is based on the following logarithmic counterpart of Lemma 3.1, where the use of the isoperimetric inequality is replaced with Lemma 6.1.
Lemma 6.3.
Suppose that is in and solves the heterogeneous distortion inequality (1.1) with , and . If is bounded and not the constant function , then for every and almost every ball , we have
Proof.
Proof of Theorem 1.2.
Let then . By Lemma 6.3 and Hölder’s inequality, we have
Since this holds for every and a.e. , Lemma 3.2 yields that
where and are independent of our choice of . Hence, by Lemma 3.3, we have that is Hölder continuous in . Therefore, the function locally Hölder continuous in . In particular, is locally bounded. However, if for some , then . We conclude that cannot have any zeroes. ∎
7. The Liouville theorem
The remaining part of this paper is devoted to proving the Liouville theorem formulated in Theorem 1.3.
We recall from the introduction that our approach is to consider a function “” from to . This mapping is well defined and satisfies a similar distortion inequality as the classical complex logarithm map. The differential inequality makes it possible to show that the weak derivative of our “” lies in for some . The argument for this goes back to two remarkable papers by Iwaniec and Martin [18] (for even dimensions) and Iwaniec [16] (for all dimensions), where they proved local integral estimates of quasiregular mappings below the natural exponent . Later, a short proof was given by Faraco and Zhong [8]. We in turn perform a global version of the Lipschitz truncation argument of Faraco and Zhong in our setting.
7.1. The logarithm with a manifold target
Let . Then there exists a smooth mapping , defined by
for and . The map is conformal, with . The inverse of is given by
for . A simple calculation yields that , and therefore
We use the inverse to take a “logarithm” of our mapping .
Lemma 7.1.
Suppose that satisfies the heterogeneous distortion inequality (1.1) with and , for some . Suppose also that is bounded, and that is not the constant mapping . Denote
Then has the following properties:
-
(1)
is continuous and ;
-
(2)
we have ;
-
(3)
we have
for a.e. .
7.2. Integrability below the natural exponent
According to Lemma 7.1, the logarithmic mapping lies in and it solves the distortion inequality,
(7.1) |
Since , the integral of the Jacobian over vanishes. Therefore, the natural integral estimate for the logarithmic map over the entire space reads as follows
The next lemma gives the key global integrability estimate for the differential below the natural exponent .
Lemma 7.2.
Suppose that a mapping is continuous, and that with . If satisfies the distortion inequality (7.1) with for some , then there exists such that . In particular, we have the estimate
where as our choice of tends to .
Proof.
We may assume . We denote , where and . Let
and for every , let
Suppose that . Then by a pointwise Sobolev estimate, we have for every that
Hence, is -Lipschitz in . Consequently, by using the McShane extension theorem [23], we find a -Lipschitz map such that . We denote .
We point out that we have
a.e. in . Indeed, we have a.e. in , and since , we also have a.e. in . Since , we also have , and therefore . Hence, we may apply the case of Lemma 2.4 with a manifold target, obtaining that
For , we denote . Since in , we may therefore estimate that
Moreover, since , we have
We now chain these estimates together, and multiply by , where . We obtain
(7.2) |
Let . We now integrate (7.2) from to with respect to , and use the Fubini–Tonelli theorem to switch the order of integration. Observe that if , and otherwise. Hence,
for a.e. . Moreover, we also have
and hence
for a.e. . In conclusion, we obtain the estimate
(7.3) |
We then further estimate some of the terms in (7.3). On the left hand side, we observe that if , then , and therefore . Hence, we have
For the first term on the right hand side, we see from the definition of that , and therefore obtain
For the remaining terms, we use the strong Hardy–Littlewood maximal inequality, where the same constant can be used for all exponents in the interval . Moreover, we also estimate the third term by After all these estimates of individual terms, we obtain a total estimate of the form
(7.4) |
We then use Young’s inequality to obtain the estimate
Hence, combining this with (7.4), we now have
(7.5) |
where . We then select small enough that . Since is of finite measure, is integrable over it, and we may absorb its term from the right hand side of (7.5) to the left hand side. We obtain the estimate
We let . Since , this makes the final term vanish, yielding
Note that we may assume that . Indeed, otherwise has a zero; this is possible only if , in which case is constant and the claim is trivial. Hence, by letting , the claim follows. ∎
7.3. Proof of the Liouville theorem
It remains to complete the proof of Theorem 1.3. We recap the statement before the proof.
Theorem 1.3.
Suppose that satisfies the heterogeneous distortion inequality (1.1) with and , for some . If is bounded and , then .
Proof.
Suppose towards to a contradiction that is bounded and , but is not identically zero. By Theorems 1.1 and 1.2, we have that is continuous and has no zeros. Hence, we may define the “logarithmic” mapping by
By Lemma 7.1, we have that , , and . Combining this with Lemma 7.2 we conclude that for some . In particular, since , we have
(7.6) |
Consider now balls of the form . Our goal is to show that the integral average of over , denoted by , is bounded independently of . By the Sobolev-Poincaré inequality [7, 4.5.2] and (7.6) we have
Consequently, we have that
Since by Lemma 5.4, we have . However, since , we have . This leads to a contradiction, and the claim follows. ∎
References
- [1] K. Astala, T. Iwaniec, and G. Martin. Elliptic partial differential equations and quasiconformal mappings in the plane. Princeton university press, 2009.
- [2] K. Astala and L. Päivärinta. Calderón’s inverse conductivity problem in the plane. Ann. of Math. (2), 163(1):265–299, 2006.
- [3] R. Caccioppoli. Limitazioni integrali per le soluzioni di un’equazione lineare ellitica a derivate parziali. Giorn. Mat. Battaglini (4), 4(80):186–212, 1951.
- [4] S. Campanato. Equazioni ellittiche del ordine espazi . Ann. Mat. Pura Appl. (4), 69:321–381, 1965.
- [5] A. Cauchy. C.R. Acad. Sci. Paris, (19):1377–1384, (1844).
- [6] A. Convent and J. Van Schaftingen. Intrinsic co-local weak derivatives and Sobolev spaces between manifolds. Ann. Sc. Norm. Super. Pisa Cl. Sci. (5), 16(1):97–128, 2016.
- [7] L. C. Evans and R. F. Gariepy. Measure theory and fine properties of functions. Textbooks in Mathematics. CRC Press, Boca Raton, FL, revised edition, 2015.
- [8] D. Faraco and X. Zhong. A short proof of the self-imroving regularity of quasiregular mappings. Proc. Amer. Math. Soc., 134(1):187–192, 2005.
- [9] I. Fonseca and W. Gangbo. Degree theory in analysis and applications, volume 2 of Oxford Lecture Series in Mathematics and its Applications. The Clarendon Press, Oxford University Press, New York, 1995. Oxford Science Publications.
- [10] F. W. Gehring. The -integrability of the partial derivatives of a quasiconformal mapping. Acta Math., 130:265–277, 1973.
- [11] P. Hajłasz, T. Iwaniec, J. Malỳ, and J. Onninen. Weakly differentiable mappings between manifolds. Mem. Amer. Math. Soc., 192(899), 2008.
- [12] L. I. Hedberg. On certain convolution inequalities. Proc. Amer. Math. Soc., 36:505–510, 1972.
- [13] J. Heinonen, T. Kilpeläinen, and O. Martio. Nonlinear potential theory of degenerate elliptic equations. Dover, 2006.
- [14] S. Hencl and P. Koskela. Lectures on mappings of finite distortion, volume 2096 of Lecture Notes in Mathematics. Springer, Cham, 2014.
- [15] S. Hencl and K. Rajala. Optimal assumptions for discreteness. Arch. Ration. Mech. Anal., 207(3):775–783, 2013.
- [16] T. Iwaniec. -harmonic tensors and quasiregular mappings. Ann. of Math., 136(3):589–624, 1992.
- [17] T. Iwaniec, P. Koskela, and J. Onninen. Mappings of finite distortion: monotonicity and continuity. Invent. Math., 144(3):507–531, 2001.
- [18] T. Iwaniec and G. Martin. Quasiregular mappings in even dimensions. Acta Math., 170(1):29–81, 1993.
- [19] T. Iwaniec and G. Martin. Geometric function theory and non-linear analysis. Clarendon Press, 2001.
- [20] T. Iwaniec, C. Scott, and B. Stroffolini. Nonlinear Hodge theory on manifolds with boundary. Ann. Mat. Pura. Appl. (4), 177(1):37–115, 1999.
- [21] T. Iwaniec and V. Šverák. On mappings with integrable dilatation. Proc. Amer. Math. Soc., 118(1):181–188, 1993.
- [22] R. Mañé, P. Sad, and D. Sullivan. On the dynamics of rational maps. Ann. Sci. École Norm. Sup. (4), 16(2):193–217, 1983.
- [23] E. J. McShane. Extension of range of functions. Bull. Amer. Math. Soc., 40(12):837–842, 1934.
- [24] C. B. Morrey, Jr. On the solutions of quasi-linear elliptic partial differential equations. Trans. Amer. Math. Soc., 43(1):126–166, 1938.
- [25] C. B. Morrey, Jr. Second-order elliptic systems of differential equations. In Contributions to the theory of partial differential equations, Annals of Mathematics Studies, no. 33, pages 101–159. Princeton University Press, Princeton, N. J., 1954.
- [26] C. B. Morrey, Jr. Multiple integrals in the calculus of variations. Die Grundlehren der mathematischen Wissenschaften, Band 130. Springer-Verlag New York, Inc., New York, 1966.
- [27] J. Onninen and X. Zhong. Mappings of finite distortion: a new proof for discreteness and openness. Proc. Roy. Soc. Edinburgh Sect. A, 138(5), 2008.
- [28] Y. G. Reshetnyak. Bounds on moduli of continuity for certain mappings. Sibirsk. Mat. Zh., 7:1106–1114, 1966. (Russian).
- [29] Y. G. Reshetnyak. The Liouville theorem with mininal regularity conditions. Sibirsk. Mat. Zh., 8:835–840, 1967. (Russian).
- [30] Y. G. Reshetnyak. On the condition of the boundedness of index for mappings with bounded distortion. Sibirsk. Mat. Zh., 9:368–374, 1967. (Russian).
- [31] Y. G. Reshetnyak. Space mappings with bounded distortion. Sibirsk. Mat. Zh., 8:629–659, 1967. (Russian).
- [32] Y. G. Reshetnyak. Space mappings with bounded distortion, volume 73 of Translations of Mathematical Monographs. American Mathematical Society, Providence, RI, 1989.
- [33] S. Rickman. Quasiregular mappings, volume 26. Springer-Verlag, 1993.
- [34] E. Villamor and J. J. Manfredi. An extension of Reshetnyak’s theorem. Indiana Univ. Math. J., 47(3):1131–1145, 1998.
- [35] S. K. Vodop’janov and V. M. Gol’dšteĭn. Quasiconformal mappings, and spaces of functions with first generalized derivatives. Sibirsk. Mat. Ž., 17(3):515–531, 715, 1976.
- [36] W. P. Ziemer. Weakly differentiable functions, volume 120 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1989. Sobolev spaces and functions of bounded variation.
- [37] V. A. Zorič. M. A. Lavrent’ev’s theorem on quasiconformal space maps. Mat. Sb. (N.S.), 74 (116):417–433, 1967.