Removable Singularities of Harmonic Functions on Stratified Sets (2024)

1. Introduction

There are deep historical connections between symmetry, harmonic functions, and stratified sets. A beautiful paper [1] by G. W. Mackey gives a systematic overview of the interplay of harmonic analysis and symmetry. At the same time, stratified sets often appear as quotients under symmetry group actions (see, for instance, [2]).

In 1926, R. Courant considered (see [3]) the oscillations of an elastic medium of a composite type: a membrane with an attached grid of strings. After a timelapse, at the end of the 1950s, M. Schechter published a series of articles that can be attributed to this subject too (see, for instance, [4]). Systematic studies in the field began at the end of the 1980s and were related to vibrations of string systems that have the form of a one-dimensional stratified set (in other words, a geometric graph, quantum graph, or topological network; see, for instance, [5]). Finally, in the 1990s, the first generalizing attempts were made to construct the theory of differential equations describing the oscillations or equilibria of intricate structures composed of elastic continua of arbitrary dimension that adjoin each other like cells in a cell complex. Such a structure is called a stratified set.

Significant progress was made in the field of elliptic second-order partial differential equations on stratified sets. The progress was achieved due to new principles for modeling such structures based on the concept of stratified measure and differential operators induced by this measure (see [6]).

The description of removable singularities for solutions of partial differential equations in a given functional space traditionally attracts significant attention from researchers. The classical result in this direction is the theorem claiming the removability of a relatively closed set of zero (harmonic) capacity for a bounded harmonic function on a domain of Euclidean space Rn.

In this paper, we prove an analog of this theorem for bounded harmonic functions in the sense of the “soft Laplacian” on stratified sets with flat interior strata.

The main result asserts that for a bounded harmonic function on an n-dimensional stratified set satisfying the “strong sturdiness condition”, a relatively closed set is removable whenever its intersection with the closure of any n-dimensional stratum has capacity zero in that stratum.

This result can become one of the main technical components for extending the well-known Poincaré–Perron’s method of proving the solvability of the Dirichlet problem for a soft Laplacian. Previously (see [7]), this could only be done in the two-dimensional case.

This paper is further organized as follows: In Section 2, we collect preliminary information, including the definitions of a stratified set (Section 2.2); stratified measure (Section 2.2); and basic differential operators: gradient, divergence, and Laplacian (Section 2.3). In Section 2.4, we recall the mean value theorem and Harnack’s inequality for harmonic functions on stratified sets.

In Section 3, we establish our main result, namely, the removable singularity theorem, which is a straightforward consequence of Lemmas 1 and 2. The proof of Lemma 2 relies on Lemma 3, which has a rather long technical proof.

Section 4 is entirely devoted to the proof of Lemma 3. In Section 4.1, we present estimates for the gradient of a harmonic function on a stratified set (Theorem 4). These estimates generalize the corresponding results for harmonic functions in Euclidean space. In Section 4.3, we prove Theorem 5, claiming that the gradient of a harmonic function has zero flux through any admissible sphere. For harmonic functions in Euclidean space, this fact is a direct consequence of the divergence theorem, but in the setting of stratified sets, its proof becomes delicate in view of possible unboundedness of the gradient of a harmonic function near strata of dimension n2, where n is the dimension of the stratified set under consideration. The proof of Lemma 3 is finished in Section 4.3, and this completes the proof of the main result.

In the closing Section 5, we formulate the conclusions of the research.

2. Preliminaries

In this section we collect preliminary information concerning stratified sets, basic differential operators, and harmonic functions on stratified sets. In a more general setting, the definitions for stratified set and related notions can be found in [8,9]. To a great extent, these definitions are inspired by [10].

2.1. StratifiedSets

In this paper, by a stratified setΩ, we mean a connected subset of the Euclidean space RN that consists of finitely many pairwise disjoint connected (boundaryless) submanifolds called strata. The set of all strata is denoted by Σ, while the strata themselves are denoted by σkj:

Ω=σkjΣσkj.

The first subscript indicates the dimension of a stratum, while the second enumerates the strata of the given dimension. We impose the following requirements on the mutual disposition of strata:

  • The closure σ¯kj of every stratum is compact and the boundary σkj=σ¯kjσkj is the union of some strata in Σ;

  • For any two strata σkj,σmiΣ, the intersection of their closures σ¯kjσ¯mi is either empty or consists of some strata in Σ.

Henceforth, the relation σkjσmi designates that σkjσmi. In this case, we say that the strata are contiguous (to one another).

We use the interior metricd on Ω, where d(X,Y) is defined as usual to be the infimum of the lengths of curves through X,YΩ in Ω. It is easy to see that the metric topology agrees with the topology induced in Ω by the inclusion ΩRN. All topological notions below refer to this topology.

Also, we represent Ω as the union Ω°Ω° (“interior” and “boundary”), in which Ω° is an open connected subset of Ω composed of some strata in Σ and satisfies the equality Ω°¯=Ω; the remaining part Ω°=ΩΩ° is then the topological boundary of Ω°.

In this paper, all interior strata are assumed to be flat in the following sense: every stratum σkjΩ° is a subdomain of a k-dimensional affine subspace of RN.

2.2. StratifiedMeasure

We call a set ωΩμ-measurable if every intersection σkjω is measurable with respect to the k-dimensional Lebesgue measure on σkj. It is easy to see that the set MΩ of all μ-measurable sets is a σ-algebra on Ω. The stratified measureμΩ on Ω (more precisely on MΩ) is defined as

μΩ(ω)=σkjΣμk(ωkj),

where μk(ωkj) is the k-dimensional Lebesgue measure of the set ωkj=ωσkj. The measurability of a function f:ΩR is defined routinely: f is μ-measurable if all Lebesgue sets Lf(c)={XΩ:f(X)c} belong to MΩ for cR. It is easy to see that the Lebesgue integral of a μ-measurable function over a μ-measurable set ω reduces to the sum

ωfdμΩ=σkjΣωkjfdμk.

We frequently omit Ω in the notation μΩ, hoping that the stratified set under consideration is uniquely identified by the context.

2.3. Gradient, Divergence, andLaplacian

Henceforth, by C1(Ω°), we mean the space of tangent vector fields F on Ω° whose restrictions F|σki to the strata σkiΩ° belong to the spaces C1(σkj).

The divergence of a vector field FC1(Ω) at a point XσkjΩ° is defined to be

·F(X)=k·F(X)+σk+1iσkjF(X+0·νi)·νi,

where the summation is carried out over all (k+1)-dimensional strata σk+1i contiguous to σkj. Here, k on the right-hand side denotes the operator of the conventional k-dimensional divergence applied to the restriction F|kj of F to σkj, νi is the unit inward normal to σkj in σk+1j at X, and F(X+0·νi) is the limit of F(Y) as Yσk+1i tends to X inside σk+1iσkj in the direction of νi. See Figure 1 for an illustration.

The so-defined divergence is a genuine analog of the classical one. It can be shown that as in the ordinary setup, the divergence ·F(X) is the density of the flux of the vector field F at X with respect to the stratified measure μ defined in the previous section. Furthermore, for FC1(Ω°), we have the following stratified analog of the divergence theorem (see [6]; the sign “−” is caused by the choice of the inward normal):

Ω°·Fdμ=Ω°Fνdμ,

where for XσkjΩ°:

Fν(X)=F(X+0·νi)·νi

with the summation taken over all σk+1iσkj not lying in Ω°.

For a sufficiently smooth scalar function u, its gradient u is a tangent vector field (in this case, u is simply the collection of the gradients of the restrictions of u to the strata). It would be rather natural to define the Laplacian on a stratified set as Δu=·(u). The so-defined Laplacian is often called “hard”. At present, the qualitative theory of harmonic functions in the sense of this Laplacian is developed weakly. Therefore, here, we restrict ourselves to considering the so-called “soft” Laplacian.

We call a stratum σkjfree if it is not contiguous to any stratum of greater dimension. The soft Laplacian of a function u on Ω° is defined to be

Δ˜u=·(pu),

where p=1 on the free strata and p=0 on the remaining strata.

The explicit expression of the soft Laplacian at the points of free strata coincides with the ordinary Laplacian (in the case of nonflat strata, with the Laplace–Beltrami operator).

If the stratum σkj is not free but there exist free strata σk+1iσkj, then at a point Xσkj, the expression of the soft Laplacian looks like

Δ˜u(X)=σk+1iσkju(X+0·νi)·νi,

where the summation is taken over all free strata σk+1iσkj.

Finally, if a stratum is neither free nor contiguous to any free stratum of dimension greater by one, then, in accordance with (2), on this stratum, we have Δ˜u=0.

For an open set UΩ°, we denote C˜loc2(U) as the set of functions u:UR that satisfy the following conditions:

  • u is continuous on U;

  • For every free stratum σni, the restriction u|Uσni is twice continuously differentiable and the gradient u of the restriction has a continuous extension to each point XUσn1j of any interior stratum σn1j contiguous to σni.

A function u:UR is said to be harmonic on U if uC˜loc2(U) and u satisfies the equation

Δ˜u(X)=0

for all XU.

2.4. Mean Value Theorem and Harnack’sInequality

The functions on a stratified set that are harmonic in the sense of the soft Laplacian inherit a number of important properties of the ordinary harmonic functions. In particular, analogs of the mean value theorem and Harnack’s inequality are valid.

We call a ball Br(X0)={XΩ:d(X,X0)<r}admissible, or, in more detail, an open ball of admissible radiusr>0 with center X0, if r is less than the distance from X0 to any stratum whose closure does not contain X0. In this event, the set Sr(X0)={XΩ:d(X,X0)=r} is called an admissible sphere. Figure 2 shows several examples of admissible balls.

Admissible balls and spheres obtain a natural stratification from Ω. For example, for an admissible sphere S, all nonempty intersections σkjS, σkjΣ, are its (k1)-dimensional strata. The stratified measure μS on a stratified sphere S is defined as for any other stratified set (see Section 2.2).

Let Ω be a stratified set in which all free strata have the same dimension n. Then p=1 on all n-dimensional strata and p vanishes on all other strata in Σ.

For X0Ω° and an admissible sphere Sr(X0), consider the spherical mean:

M[Sr(X0)]u=1|Sr(X0)|pSr(X0)pudμS,

where |Sr(X0)|p=Sr(X0)pdμS.

Theorem1([11]).

(Mean Value Theorem) Let Ω be a stratified set whose free strata have the same dimension; let u be a harmonic function on Ω°; and let Sr(X0), X0Ω°, be an admissible sphere. Then:

M[Sr(X0)]u=u(X0).

Remark1.

A similar assertion is true with the means calculated over admissible balls in place of admissible spheres.

Theorem2([12]).

(Harnack’s Inequality) Let Ω be a stratified set and let K be an arbitrary compact set in Ω°. Then, the inequality

supXKu(X)CinfXKu(X)

holds for every nonnegative harmonic function u on Ω° with a constant C=C(K,Ω°) independent of u.

3. The MainResult

In this section, we establish our main result, namely, the removable singularity theorem. It is a straightforward consequence of Lemmas 1 and 2. The proof of Lemma 2 relies on Lemma 3, which has a rather long technical proof. The proof of Lemma 3 is postponed to Section 4.

Given a stratified set Ω, we define Σk as the union of all interior strata of dimension k and define Σk as the union of all interior strata of dimension at most k:

Σk=j:σkjΩ°σkj,Σk=l=0kΣl.

A stratified set Ω is called sturdy (see [12]) if all free strata have the same dimension n and the set Ω°Σn2 is connected. Let us call Ωstrongly sturdy if all free strata have the same dimension n and, for every XΣn2, there exists an admissible ball Br(X) such that the set Br(X)Σn2 is connected.

The following theorem is the main result of the present article.

Theorem3(Removable Singularity Theorem).

Let Ω be a strongly sturdy stratified set, SΩ° be a relatively closed set whose intersection with the closure of any free stratum has harmonic capacity zero in the affine space including the stratum, and u:Ω°SR be a bounded harmonic function. Then, u has a harmonic extension to all of Ω°.

The theorem is a straightforward consequence of the next two lemmas.

Lemma1.

Under the conditions of Theorem 3, the function u extends to a bounded harmonic function on Ω°Σn2.

Lemma2.

Let Ω be a strongly sturdy stratified set of dimension n and u:Ω°Σn2R be a bounded harmonic function. Then, u has a harmonic extension to all of Ω°.

Remark2.

The strong sturdiness condition is essential.

As an example, consider a two-dimensional set Ω composed of two planar triangles with one common vertex σ01 and no other points in common (see Figure 3). Set Ω°=, i.e., Ω°=Ω. The function u, which is equal to 0 on one triangle and to 1 on the other, is harmonic on Ωσ01, but fails to have a harmonic extension to all of Ω.

ProofofLemma1.

It suffices to prove that for every point X0ΣnΣn1 and any admissible ball Br(X0), the restriction u|Br(X0)S admits an extension to a bounded harmonic function on Br(X0).

If X0Σn, then the existence of a sought extension follows from the removability of a relatively closed set of capacity zero for bounded harmonic functions on subdomains of Rn. Therefore, suppose that X0 belongs to an (n1)-dimensional stratum σn1 and let Br(X0) be an admissible ball. Without loss of generality, we assume Ω to be the closed stratified ball Br(X0)Sr(X0) in which Br(X0) serves as the interior and Sr(X0) as the boundary (see Figure 4).

Let σn1,,σnm be all strata contiguous to σn1.

If we have exactly two strata, i.e., σn1 and σn2, then Br(X0) (with its interior metric) is isometric to the usual Euclidean ball B of radius r in Rn; moreover, σn1 and σn2 go under isometry into half-balls separated by the (n1)-dimensional disk that is the image of σn1. The function u˜ corresponding to u under isometry is an ordinary harmonic function on BS˜, where S˜ is the image of SBr(X0) under isometry. Applying the standard removable singularity theorem to u˜, we obtain a harmonic extension of u˜ to B. Executing the inverse isometry, we obtain a harmonic extension of u to Br(X0).

If we have an odd number of strata, then we double their number by slightly rotating the ball Br(X0) around σn1 and extend u to the new strata by using this rotation.

Thus, we may assume that there are 2l strata in total, i.e., σnj, j=1,,2l, that are contiguous to σn1. Again, we take a Euclidean ball B of radius r in Rn divided into two half-balls B1 and B2 resting on an equatorial (n1)-dimensional disk D. Let J{1,2,,2l} be an arbitrary collection of l distinct numbers and let J¯ be the complementary collection. Now, map σnjσn1, jJ, onto B1D and map σnj¯σn1, j¯J¯, onto B2D isometrically so that the stratum σn1 goes onto D. Let S˜ be the union of the images of (σnjσn1)S over all j under isometry. Translating u by isometry from σnjσn1 to B1D, we receive functions uj, jJ. Proceeding similarly, we also receive functions uj¯, j¯J¯. It is easily seen that the function uJ, which is equal to jJuj on (B1D)S˜ and equal to j¯J¯uj¯ on (B2D)S˜ (notice that both the functions agree on the intersection of their domains), is an ordinary harmonic function on BS˜. Applying to it the standard removable singularity theorem, we obtain a harmonic extension of uJ to the whole ball B. Since this is true for every collection J, easy combinatorics show that each individual function uj obtains an extension rather than their combination. Returning to u, we obtain an extension of u to the whole set σnjσn1 for every j; furthermore, the resultant function is harmonic on the whole ball Br(X0). □

ProofofLemma2.

Equation (3) imposes no conditions on the strata of dimension n2. Thus, the lemma asserts, in fact, that u extends continuously to Ω°.

We extend u from Ω°Σn2 to Ω° by using spherical means. To this end, we state the following lemma whose proof is postponed to the next section.

Lemma3(EqualityofMeans).

Under the conditions of Theorem 3, for every X0Ω° and any admissible radii r1 and r2, the equality of means holds:

M[Sr1(X0)]u=M[Sr2(X0)]u.

Leaning on this lemma, we set

M(X)=M[Sr(X)]u,XΩ°,

where r is any admissible radius.

Note that

u(x)=M(X)forXΩ°Σn2.

Let us prove by induction on k decreasing from k=n1 to 0 that M(X) is continuous on Ω°Σk1. Then, M(X) gives us the sought after extension of u.

The induction base is trivial. Suppose that M(X) is continuous on Ω°Σk, k<n1, and hence harmonic on Ω°Σk. Take X0σk for a k-dimensional stratum σkΩ°. We are to prove that

limXX0M(X)=M(X0).

This justifies the induction step and, therefore, completes the proof of Lemma 2.

The proof of (4) is divided into Lemmas 4–6. □

Lemma4.

limXX0,XσkM(X)=M(X0).

Proof.

Let Xiσk, i=1,2,, and XiX0 as i. Let Sr(X0) be an admissible sphere and let BR(X0) be an admissible ball of a greater radius R>r. Since the points Xi lie on the same stratum as X0, the translated spheres Sr(Xi) are admissible too and lie within BR(X0) for large enough i. For ZBR(X0)Σn2, define

ui(Z)=u(Z+(XiX0)).

The sequence ui converges pointwise to u and is uniformly bounded. Therefore,

Sr(X0)puidμSSr(X0)pudμS.

Hence,

M(Xi)=M[Sr(Xi)]u=M[Sr(X0)]uiM[Sr(X0)]u=M(X),

which finishes the proof of the lemma. □

Lemma5.

m:=lim infXX0,XΩσkM(X)=M(X0).

Proof.

Choose XiΩσk so that XiX0 and M(Xi)m. Let Lk be the k-dimensional affine plane including σk. Let Yi be the projection of Xi to Lk. Since XiX0 and YiX0, the balls Bri(Yi), ri=2|XiYi|, are admissible for large enough i (see Figure 5).

Since ri0, while m is the lower limit of M(X) at X0, it follows that

mi:=infXBri(Yi)σkM(X)masi.

Indeed, for any ε>0, there exists ZiBri(Yi)σk such that

M(Zi)ε<miM(Xi).

Then

lim infiM(Zi)εlim infimi,

lim supimilimiM(Xi)=m.

Notice that ZiX0 as i. Therefore,

lim infiM(Zi)m.

Inequalities (6) and (8) imply the inequality

mεlim infimi

which, by the arbitrariness of ε, yields

mlim infimi.

In turn, (9) and (7) imply the validity of (5).

Put

ui(X):=M(X)mi,XBri(Yi)σk.

Then, ui are nonnegative bounded functions that are harmonic on Bri(Yi)σk according to the induction assumption; moreover, ui(Xi)0 as i.

Since all Yi and X0 lie on the same stratum, all (admissible) balls Bri(Yi) are similar to one admissible ball Br(X0). Let Ti:Bri(Yi)Br(X0) denote the corresponding similarity transformation. Then,

Ti(Yi)=X0,dist(Ti(Xi),σk)=r/2,|Ti(Xi)X0|=r/2.

For ZBr(X0), define

u¯i(Z):=ui(Ti1(Z)).

Then, u¯i are nonnegative bounded harmonic functions on Br(X0)σk.

Let HBr(X0)σk be an arbitrary compact set. Since all points Ti(Xi) are at a positive distance from the boundary of Br(X0)σk, there exists a compact set KBr(X0)σk including H, as well as all the points Ti(Xi). Taking it into account that u¯i(Ti(Xi))=ui(Xi)0 as i and applying Harnack’s inequality to the functions u¯i on Br(X0)σk and the compact set K (it is here that we use the strong sturdiness condition of Ω), we conclude that the functions u¯i converge uniformly to zero on K and, as a consequence, on H.

Since the compact set H can be taken arbitrarily close to σk and the u¯i are uniformly bounded, the means of u¯i over the spheres with center X0 tend to zero. Indeed, let C be an upper bound for the functions u¯i and let ε>0 be arbitrary. For ρ>0, set

σk(ρ)={ZBr(X0):dist(Z,σk)<ρ}.

Choose ρ so small that

|Sr/2(X0)σk(ρ)|p<ε|Sr/2(X0)|p/(2C).

Now, take the compact set

H=Sr/2σk(ρ).

Choose N to have u¯i(Z)<ε/2, ZH, for i>N. Then, for i>N:

M[Sr/2(X0)u¯i]=1|Sr/2(X0)|pHp¯uidμS+Sr/2Hpu¯idμS<1|Sr/2(X0)|pε|H|p/2+C|Sr/2(X0)H|pε.

For Sri/2(Yi)=Ti1Sr/2(X0), we obtain

M[Sri/2(Yi)]ui=M[Sr/2(X0)]u¯i0(i).

Therefore, M(Yi)m as i because the spherical means of the functions M(X) and u(X) coincide in view of the equality M(X)=u(X) for XΩ°Σn2.

From Lemma 4, we also have M(Yi)M(X0). Hence, m=M(X0). □

Lemma6.

M(X0)=lim supXX0,XΩσkM(X).

For a proof, it suffices to apply Lemma 5 to the function u, leaning upon the equality lim inf(u)=lim supu.

Lemmas 5 and 6 yield limXX0M(X)=M(X0). This and Lemma 4 imply (4), thus finishing the Proof of Lemma 2.

4. Proof of Lemma 3

This section is entirely devoted to the proof of Lemma 3. In Section 4.1, we present estimates for the gradient of a harmonic function on a stratified set (Theorem 4). These estimates generalize the corresponding results for harmonic functions in Euclidean space. In Section 4.2, we prove Theorem 5, claiming that the gradient of a harmonic function has zero flux through any admissible sphere. For harmonic functions in Euclidean space, this fact is a direct consequence of the divergence theorem, but in the setting of stratified sets, its proof becomes delicate in view of possible unboundedness of the gradient of a harmonic function near strata of dimension n2, where n is the dimension of the stratified set under consideration. The proof of Lemma 3 is finished in Section 4.3, and this completes the proof of the main result.

4.1. GradientEstimate

We need an estimate for the gradient of a harmonic function on a stratified set. In the classical case when u is a bounded harmonic function on a domain GRn, we have the following gradient estimate:

|u(X)|nρ(X)supG|u|,

where ρ(X)=dist(X,G) (see, for instance, [13]). We prove a similar estimate in the stratified setup.

Theorem4(Gradient Estimate).

Let Ω be a stratified set whose free strata have the same dimension n and let u:Ω°Σn2R be a bounded harmonic function. Then, for every XΩ°Σn2, the following estimate holds:

|u(X)|Cρ(X)supΩ°Σn2|u|,

where C=C(Ω) is a constant depending only on the structure of the stratified set Ω and ρ(X)=dist(X,Σn2Ω°).

We first give an estimate over admissible balls.

Lemma7.

For every X0Ω°Σn2 and any admissible ball Br(X0), the following estimate holds:

|u(X)|CrsupBr(X0)|u|,XBr/2(X0),

where C=C(Ω) is a constant depending only on the structure of Ω.

Proof.

We use the same trick as in the proof of Lemma 1.

If X0 belongs to an n-dimensional stratum, then the required estimate follows from the above-mentioned classical result. We, therefore, assume that X0 belongs to some (n1)-dimensional stratum σn1 and let Br(X0) be an admissible ball. Without loss of generality, we consider Ω to be the closed stratified ball Br(X0)Sr(X0) in which Br(X0) serves as the interior and Sr(X0) as the boundary (Figure 4).

Let σn1,,σnm be all strata contiguous to σn1.

If we have only two strata, namely, σn1 and σn2, then Br(X0) (with its interior metric) is isometric to the ordinary Euclidean ball of radius r in Rn; moreover, under isometry, the strata σn1 and σn2 go onto two half-balls separated by an (n1)-dimensional disk, which is the image of σn1. The function u˜ corresponding to u under isometry is an ordinary harmonic function on the Euclidean ball and the required estimate is valid for it. By isometry, the required estimate holds for u too.

Proceeding as in the proof of Lemma 1, we may assume that we have 2l strata σnj, j=1,,2l, contiguous to σn1. Repeating the construction from the proof of Lemma 1, we take a Euclidean ball B of radius r in Rn divided into two half-balls B1 and B2 resting on an (n1)-dimensional disk D. Next, for an arbitrary collection J{1,2,,2l} of l distinct numbers and the complementary collection J¯, we obtain the functions uj on B1D, jJ, and the functions uj¯ on B2D, j¯J¯, that correspond to the functions u|σnjσn1, jJ, and the functions u|σnj¯σn1, j¯J¯, by isometry. Then, the function uJ, which is equal to jJuj on B1D and equal to j¯J¯uj¯ on B2D, is an ordinary harmonic function on B, and for it, we have the required estimate. Since this is true for any J, easy combinatorics show that the sought estimate is valid for each function uj, j=1,,2l, and for the function u by isometry. □

Another proof can be accessed by utilizing the Poisson integral representation in admissible stratified balls centered at points of an (n1)-dimensional stratum.

The following lemma is geometrically quite evident.

Lemma8.

If XΣn, then dist(X,Σn1Ω°) is an admissible radius for X. There exists α(0,1](depending on Ω)such that if XΣn1, then αdist(X,Σn2Ω°) is an admissible radius for X.

ProofofTheorem4.

Take XΩ°Σn2 and

ρ=dist(X,Σn2Ω°).

If dist(X,Σn1)<αρ/4, then there is an (n1)-dimensional stratum σn1 and a point X0σn1 such that XBαρ/4(X0). Observe that dist(X0,Σn2Ω°)>ρ/2. By Lemma 8, αρ/2 is an admissible radius for X0, and by Lemma 7,

|u(X)|2CαρsupBαρ/2(X0)|u|.

If dist(X,Σn1)αρ/4, then, by Lemma 8, αρ/4 is an admissible radius for X, and by Lemma 7,

|u(X)|4CαρsupBαρ/4(X)|u|.

Either way, we obtain the required estimate. □

4.2. GradientFlux

Theorem5.

Let Ω be a stratified set whose free strata have the same dimension n and let u: Ω°Σn2R be a bounded harmonic function. Then, u has zero flux through any admissible sphere Sr(X):

Sr(X)(pu)νdμS=0.

Proof.

If Br(X)Σn2=, then the conclusion of the lemma follows from the application of the divergence theorem, which is expressed by (1), to the vector field puC1(Br(X)), with the equality ·(pu)=0 taken into account.

Therefore, assume Br(X)Σn2. Let

σd11,σd21,,σd2j2,,σdk1,,σdkjk

be all strata whose closures contain X and dimensions do not exceed n2. Consider the dimensions to increase:

d1<d2<<dkn2.

Note that there is exactly one stratum, namely, σd11, of dimension d1, and Xσd11.

Let 0<ρ1<r and

σd11(ρ1)={ZBr(X):dist(Z,σd11)<ρ1}.

Choose ρ2, 0<ρ2<ρ1, so that the sets

σd2j(ρ2)={ZBr(X):dist(Z,σd2j)<ρ2}σd11(ρ1)

be pairwise disjoint for j=1,,j2 and do not intersect the strata σdij, i>2, that are not contiguous to σd2j.

Proceeding by induction, finally choose ρk, 0<ρk<ρk1, so that the sets

σdkj(ρk)={ZBr(X):dist(Z,σdkj)<ρk}ik1,sσdis(ρi)

are pairwise disjoint for j=1,,jk.

Set for l=1,,k

Σ(ρ1,,ρl)=il,sσdis(ρi)

and consider the stratified set

Pl=Br(X)Sr(X)Σ(ρ1,,ρl)

with boundary

Pl°=Sr(X)Σ(ρ1,,ρl)il,sΓdisl,

where

Γdisl=Γdisl(ρ1,,ρl)={ZBr(X):dist(Z,σdis)=ρi}Σ(ρ1,,ρl).

Figure 6 shows what remains after erasing the “cylinders” σdkj(ρk) from the ball; in three-dimensional space, the process of constructing Pl is not very diverse and might involve only one or two steps that depend on the dimension of the stratum containing the center of the ball. In this figure, the invisible part of the sphere is also included in the boundary Pl°.

Inducting on l decreasing from l=k to 0, let us prove that

Pl°(pu)νdμ=0,

where dμ is the stratified measure on Pl°.

Take l=k. The set Σ(ρ1,,ρk) represents a neighborhood of the set Σn2Br(X). Then, puC1(Pk°), and the divergence theorem (see (1)) gives

Pk°(pu)νdμ=0,

which justifies the induction base.

Assuming (11) is true, let us validate the equality for l1. We have

Pl°=Pl1°sσdls(ρl)sΓdlsl.

Hence,

Pl1°s=1jlσdls(ρl)(pu)νdμ=s=1jlΓdlsl(pu)νdμ,

where the signs of the integrals on the right-hand side are chosen in accordance with the fact that Γdlsl is part of the boundary of σdls(ρl).

The next relations are justified below:

limρl0Pl1°s=1jlσdls(ρl)(pu)νdμ=Pl1°(pu)νdμ,

limρl0Γdlsl(ρ1,,ρl)(pu)νdμ=0.

Applying them, we obtain the following from (12):

Pl1°(pu)νdμ=0,

completing the validation of the induction step.

Since P0°=Sr(X), the equality (11) for l=0 coincides with (10).

To finish the proof of the theorem, it remains to justify (13) and (14).

Since p=1 on the strata of dimension n and p=0 on the other strata, we have

Pl1°(pu)νdμ=jσnjPl1°(u)νdμn1,

where the summation is over all n-dimensional strata whose closures contain X, and dμn1 is the surface (n1)-measure.

By Theorem 4, the following estimate holds:

|u(Z)|Cdist(Z,Σn2).

The following lemma is rather obvious and we leave it without proof.

Lemma9.

Let Ld be a d-dimensional affine subspace of Rn, dn2, and G be an (n1)-dimensional compact piecewise smooth surface in Rn, which is smooth in a neighborhood of the intersection GLd and is transversal to Ld. Then, the function

f(x)=1dist(x,Ld)

is integrable with respect to the surface (n1)-measure on G.

Note that for Zσnj:

1dist(Z,Σn2)i,s1dist(Z,Ldis),

where Ldis is the affine subspace including σdis, and dist(Z,Ldis) is the distance from Z to Ldis inside Lnj. From (15) and (16) and Lemma 9, we then conclude that (pu)ν is μ-integrable on Pl1°. This proves (13).

By analogy to (15), we have

Γdlsl(ρ1,,ρl)(pu)νdμ=jσnjΓdlsl(u)νdμn1,

with the summation taken over the n-dimensional strata contiguous to σdls.

Fix a stratum σnj contiguous to σdls and consider

φ(ρl)=σnjΓdlsl(ρ1,,ρl)(u)νdμn1=σnjΓdlsl(ρ1,,ρl)uνdμn1.

First, suppose that dln3. In view of (16),

|φ(ρl)|σnjΓdlsl(ρ1,,ρl)Cdist(Z,Σn2)dμn1.

The following assertion is proven by straightforward calculations.

Lemma10.

Let LdLm be a d-dimensional and an m-dimensional affine subspaces of Rn, dmn2, let σdLd be a bounded open subset, and

Γ(ρ)={xRn:dist(x,σd)=ρ}.

Then,

Γ(ρ)1dist(x,Lm)dμn1=C(d,m,n)ρnd2μd(σd).

From (19) and Lemma 10, we obtain limρl0φ(ρl)=0 when dln3.

Now, we examine the case dl=n2. In this case, l=k and dk=n2. To simplify the notations, we henceforth denote ρk by ρ, the stratum σnj by σn, the stratum σdks by σn2, and Γdksk(ρ1,,ρk) by Γ(ρ). Then,

Γ(ρ)={ZσnBr(X):dist(Z,σn2)=ρ}Σ(ρ1,,ρk1)

and Σ(ρ1,,ρk1) is a neighborhood of σn2Br(X).

Let us show that the limit limρ0φ(ρ) exists. Fix an arbitrary admissible value ρ¯>0 for ρk and, considering ρ(0,ρ¯), define the stratified set

Q={ZσnBr(X):ρdist(Z,σn2)ρ¯}Σ(ρ1,,ρk1),

whose interior Q° and boundary Q are determined relative to the topology of Ω. Note that puC1(Q), and by the divergence theorem:

Q(pu)νdμ=0.

Hence,

φ(ρ)=Γ(ρ)uνdμn1=QΓ(ρ)(pu)νdμ.

The last integral can be treated in the same way as in the deduction of (13). This proves the existence of the limit limρ0φ(ρ).

Suppose this limit is nonzero, say limρ0φ(ρ)>0.

Denote by G the projection of Γ(ρ¯) to σn2,

Γ˜(ρ)={ZσnBr(X):dist(Z,G)=ρ}

and

φ˜(ρ)=Γ˜(ρ)uνdμn1.

If ρ¯ is small enough, the (n2)-measure of the difference between G and the projection of Γ(ρ) to σn2 is arbitrarily small for ρ(0,ρ¯). In view of Theorem 4 and Lemma 10 applied with d=m=n2, we conclude that for sufficiently small ρ¯>0, the values of the functions φ(ρ) and φ˜(ρ) differ by a quantity that is arbitrarily small for ρ(0,ρ¯) and, as a consequence, the following inequality holds (with some δ):

φ˜(ρ)δ>0,ρ(0,ρ¯).

Look at the integral

0ρ¯φ˜(ρ)ρdρ=0ρ¯1ρΓ˜(ρ)uνdμn1dρ.

In the n-dimensional affine subspace Ln including σn, consider the coordinates (z,ρ,θ), where z are the Cartesian coordinates of the projection Z of ZLn to the affine (n2)-dimensional subspace Ln2 including σn2, and (ρ,θ) are the polar coordinates with center Z in the plane orthogonal to Ln2 in Ln.

Since

dμn1=ρdzdθ,

we have

0ρ¯1ρΓ˜(ρ)uνdμn1dρ=0ρ¯GΦuρ(z,ρ,θ)dθdzdρ=GΦu(z,ρ¯,θ)u(z,0,θ)dθdzCsupBr(X)Σn2|u|,

i.e., the integral is finite. On the other hand, by (20), the integral must diverge. This contradiction completes the proof of relation (14), as well as the proof of Theorem 5. □

4.3. Proof of Lemma 3

Without loss of generality, we assume Ω to have the structure of a stratified ball of admissible radius with center at a point X0. Consider spherical coordinates (r,ϕ) with center X0. Then,

1rn1dμn1=dϕ.

For small enough δ>0, we have

M[Sr+δ(X0)]uM[Sr(X0)]uδ=1ωΦu(r+δ,ϕ)u(r,ϕ)δdϕ,

where Φ is the set of directions corresponding to the (n1)-dimensional regions of the stratified sphere Sr(X0). Since u is smooth on the n-dimensional strata, it follows that as δ0:

u(r+δ,ϕ)u(r,ϕ)δu(r,ϕ)r.

By the finite increment formula, we have for some r(r,r+δ):

u(r+δ,ϕ)u(r,ϕ)δ=u(r,ϕ)r|u(r,ϕ)|.

By Theorem 4:

|u(r,ϕ)|Cdist(X(r,ϕ),Σn2Ω°)supΩ°Σn2|u|Cdist(X(r,ϕ),Σn2Ω°)supΩ°Σn2|u|.

The last inequality holds because the distance from X to Σn2 increases when X moves from X0 along a radius. Thus,

u(r+δ,ϕ)u(r,ϕ)δ

is dominated by an integrable function. By Lebesgue’s dominated convergence theorem:

M[Sr+δ(X)]uM[Sr(X)]uδ1ωΦu(r,ϕ)rdϕ=1ωrn1Sr(X)u(r,ϕ)νdμn1.

The last integral vanishes in virtue of Theorem 5.

Thus,

ddrM[Sr(X)]=0

and the value of M[Sr(X)] does not depend on r, which completes the proof of Lemma 3.

5. Conclusions

The solutions of elliptic second-order partial differential equations on stratified sets inherit many properties of solutions to such equations on domains of Euclidean space. In particular, previous investigations show that analogs of the maximum priciple, the normal derivative lemma, Harnack’s inequality, etc., are valid. In this research, we established that for harmonic functions on stratified sets, an analog of the removable singularity theorem is valid as well. This result can become one of the main technical components for extending the well-known Poincaré–Perron’s method of proving the solvability of the Dirichlet problem for a soft Laplacian.

Removable Singularities of Harmonic Functions on Stratified Sets (2024)

References

Top Articles
Latest Posts
Article information

Author: Duncan Muller

Last Updated:

Views: 5642

Rating: 4.9 / 5 (79 voted)

Reviews: 86% of readers found this page helpful

Author information

Name: Duncan Muller

Birthday: 1997-01-13

Address: Apt. 505 914 Phillip Crossroad, O'Konborough, NV 62411

Phone: +8555305800947

Job: Construction Agent

Hobby: Shopping, Table tennis, Snowboarding, Rafting, Motor sports, Homebrewing, Taxidermy

Introduction: My name is Duncan Muller, I am a enchanting, good, gentle, modern, tasty, nice, elegant person who loves writing and wants to share my knowledge and understanding with you.