Abstract
In this paper, we first make and discuss a conjecture concerning Newtonian potentials in Euclidean n space which have all their mass on the unit sphere about the origin and are normalized to be one at the origin. The conjecture essentially divides these potentials into subclasses whose criteria for membership is that a given member has its maximum on the closed unit ball at most M and its minimum at least d. It then lists the extremal potential in each subclass, which is conjectured to solve certain extremal problems. In Theorem 1, we show the existence of these extremal potentials. In Theorem 2, we prove an integral inequality on spheres about the origin, involving so-called extremal potentials, which lends credence to the conjecture.
Keywords:
harmonic functions; Newtonian potentials; Dirichlet problem; boundary Harnack inequality; Fatou theorem MSC:
31B05; 31B10; 31B20
1. Introduction
Let be a positive integer, a point in Euclidean n space, and let denote the norm of Put when For fixed let be a positive Borel measure on with Next, let denote the Hausdorff measure and let be a non-decreasing convex function on Finally, let denote the family of potentials p satisfying
Theorems 1 and 2 in this paper are based on my efforts to prove the following conjecture:
Conjecture 1.
If there is a 1-1 map from
for which there exists a potential satisfying
Given let
If and then
The analog of Conjecture 1 is true in To briefly outline its proof, we use complex notation. Therefore, and Let U denote the class of univalent (i.e, 1-1 and analytic) functions f satisfying and for which is starlike with respect to 0 (so each line segment connecting 0 to a point in D is also contained in If then based on the fact that is also starlike with respect to 0, one can show Arg (i.e., the principal argument of f) on is non decreasing as a function of , so if then
and thus Re when
From (4) and the Poisson integral formula for , it follows (see [1]) that
where is a positive Borel measure on with Note that if f is sufficiently smooth on then
Equation (6) implies that if is an arc and lies on a ray through 0, then Dividing (4) and (5) by z and integrating, we get
where log is the principal logarithm. From (7), we see that if then in (1) Given and let denote starlike univalent functions f in U satisfying in Thus, if and then for
Using this fact, one sees that the analogue of in Conjecture 1 for is where maps onto If then D is the bounded keyhole domain described as follows. For some is the union of the arcs:
and the line segments,
In [2], we showed for fixed that maps univalently onto a convex domain containing 0, and if then is subordinate to . That is, maps into This result was, in fact, a corollary of a much more general subordination theorem for Mocanu convex univalent functions that are bounded above and below in the unit disk. Our proof used a contradiction-type argument and the Hadamard–Julia variational formulas to determine the solutions to a certain class of extremal problems. Runge’s theorem then gave subordination in the given class of Mocanu convex functions. Conjecture 1 for follows from the relationship between and as well as the properties of subordination (see for example [3]).
Let denote the point in with 1 in the i th position and zeroes elsewhere. Conjecture 1 is true in when so and . It was proved in [4]. Our proof used a maximum principle for the celebrated Baernstein * function (see [5] or [6]), defined as follows: Given introduce spherical coordinates, by Let l be a locally integrable real valued function on and set
where the supremum is taken over all Borel measurable sets with
One can show that (3) in Conjecture 1 is equivalent to
For a proof of this equivalence, see Section 9.2 in [5].
For Professor Baernstein in [7] showed that if l is subharmonic in then in (8) is subharmonic in Moreover, if
then is harmonic in In need not be subharmonic in I even when l is harmonic in Instead, we proved (see also [8]) the following maximum principle: Suppose l is subharmonic and L is harmonic in
with symmetric about the axis and non increasing on Then, either in or
To briefly outline the proof of Conjecture 1 in [4], for given the existence of satisfying (2), can be deduced from a working knowledge of such tools for harmonic functions as (a) Wiener’s criteria for solutions to the Dirichlet problem, (b) the maximum principle for harmonic functions, (c) the Riesz representation formula for superharmonic functions, and (d) invariance of the Laplacian under reflection about planes containing zero (see the proof of Theorem 1 for more elaborate details). Let and for If then from the subharmonicity of harmonicity of P in decay of both potentials at and the above maximum principle, it follows that there exists with spherical coordinates and Since and on we then obtain
From this contradiction, we conclude that in Next, using and whenever that
we get in which as mentioned in (9), implies (3).
We have not been able to prove (3) in Conjecture 1 for any other values of However, in [9], we used a mass moving method in [10] to show that if and then for
Moreover, in this paper, we prove the first part of our conjecture:
Theorem 1.
Also, we prove the following:
Theorem 2.
Given as in Theorem 1, if and then
As for the plan of this paper, in Section 2, we set the stage for the proof of Theorems 1 and 2 by stating and/or proving several definitions and lemmas. In Section 3, we prove Theorem 1. In Section 4, we prove Proposition 1, a rather tedious calculation of mixed partials for a certain function. In Section 5, Proposition 1 is used to prove Theorem 2. After each theorem, we make remarks and queries.
2. Notation, Definitions, and Basic Lemmas
Throughout this paper, denotes a positive constant depending only on and Also, means is bounded above and below by positive constants whose dependence will be stated. As in Section 1, let denote the Lebesgue measure on the closure of the distance from x to the set the point in with 1 in the i th position and zeroes elsewhere, the inner product in , and the Hausdorff k measure, in
Definition 1.
If O is an open set in and is compact, then the Newtonian capacity of denoted is defined to be
where denotes the gradient of ϕ and the infimum is taken over all with on
Remark 1.
Recall that a bounded open set is said to be a Dirichlet domain if for any continuous real valued function q defined on there exists a harmonic function Q in G with
Wiener’s criteria for a bounded open set G to be a Dirichlet domain states: If
then G is a Dirichlet domain.
Given let (as in Theorem 1)
Put For one can show (see [11]) that for
Using Wiener’s criteria and the boundary maximum principle for harmonic functions, it follows that given , there exists harmonic in with continuous boundary values on and on Using the boundary maximum principle for harmonic functions, we find that if then From this fact, we deduce that uniformly on compact subsets of and is harmonic in with continuous boundary values 1 on and 0 on . Also, as and is superharmonic in an open set containing and it is, subharmonic in an open set containing From these observations and the Riesz representation theorem for sub-super harmonic functions (see [11]), it follows that there exists finite positive Borel measures with the support of , while the support of is contained in Moreover,
(once again, ).
Next, for let
and note that has boundary values that are symmetric about the axis. Thus, from the boundary maximum principle for harmonic functions and invariance of the Laplacian under rotations, we have for whenever Using this fact, and arguing as in Section 2 of [9], we get the Lebesgue–Stieltjes integral:
Here,
with
and is chosen so that With this notation, we prove the following:
Lemma 1.
Given and on when and these inequalities also hold when
This lemma is essentially trivial in but perhaps not so obvious in so we give some details.
Proof.
To begin the proof of Lemma 1, let be a plane containing the origin with unit normal, satisfying for some Let
Clearly, contains and contains If , let be the reflection of defined by We claim that if
To prove our claim, we assume, as we may, that which is permissible since are symmetric about the axis. Then, for some and We note that if then , where are orthogonal unit vectors with and are any real numbers with and Also, and Using this notation, we see that with strict inequality unless Thus, (18) is true. (18) is proved similarly, so we omit the details. To finish the proof of Lemma 1, we observe from (18) the continuous boundary values of and Harnack’s inequality for positive harmonic functions that either or is symmetric about However, this cannot happen if as is not even symmetric about the line through the origin and Similarly, in Lemma 1 now follows from these inequalities and the Hopf boundary maximum principle for harmonic functions (see [12]). □
Finally, in this section, we state
Lemma 2.
with
Proof.
We note that
where is harmonic in with continuous boundary value 0 on and as
Lemma 2 follows from (19) and essentially a boundary Harnack inequality in [13], even though is not an NTA domain for any To give a few details, if and then in is easily shown using barriers. Applying Harnack’s inequality for positive harmonic functions, we then obtain this inequality in After that, one shows for some that
Here, denotes oscillation on An iterative argument then gives Hölder continuity of in Thus, exists, and since in □
3. Proof of Theorem 1
Proof.
In the proof of existence for in Theorem 1, we assume that as existence of when was proved in [9] and if
To begin the proof of this theorem, let We assert that , where is as in Lemma 2. Indeed, since in and have continuous boundary value 0 when we see from (19) that either
To show the first possibility cannot occur, observe that it implies Also, is harmonic in with continuous boundary value on and continuous boundary value 1 on Using the fact that is the minimum and 1 the maximum of on , we can essentially repeat the proof of Lemma 1 to arrive at
Thus, is strictly decreasing on , a contradiction to (20).
Let where is chosen so that We shall show that where to complete the proof of Theorem 1, except for showing the map is 1-1. For this purpose, we note from (14) that
where is a signed measure with support in and of finite total variation. Moreover, since in with on we see that V is superharmonic in an open set containing and consequently (see [11]), is a positive Borel measure. It remains to prove that is a positive Borel measure in order to conclude the existence of in Theorem 1. For this purpose, let with Let Differentiating (22), we obtain the Poisson integral (see [11])
From the properties of the Poisson integral (see [11]) and we deduce from (23) that
for almost every where Moreover,
since V is harmonic in Also, since is smooth at points with it follows from Schauder type arguments (see [12]) that g is infinitely differentiable on From this observation and (21), we deduce that
From (24) and (26), and the mean value theorem from calculus, we arrive at
Thus, to show that it suffices to show
To prove (27), we need some boundary Harnack inequalities in [14] (see also [15]). To set the stage for these inequalities, put , where
Then, T maps
We note that
Using this note, the Kelvin transformation (see [11]), and translation invariance of harmonic functions, we find that if is harmonic at then
is harmonic at (i.e., in a neighborhood of y). From this deduction, we conclude for fixed that if then
From (28), we see that if then
Let Then, is harmonic in with continuous boundary values on and on We now are in a position to use a boundary Harnack inequality proved in Theorem 3.3 of [14] tailored to our situation. Let and set Next, put
Note that denotes the distance from y to and denotes the signed distance from to Here, denotes the boundary of relative to .
Theorem 3.
Given k, a positive integer, and there exists infinitely differentiable whenever
with for and for which as
Constants in the big O terms depend on and the norms of for on We note that in (32) have slightly different expansions since is even in while is not. From Lemma 2, (29), and (30), we find that as through coordinates of points not in
which in view of (32) implies that
Finally, showing that the map is 1-1 follows easily from the maximum principle for harmonic functions. Indeed, suppose are both mapped into Define and relative to and respectively. Let be the corresponding potentials in Theorem 1. Then, either or vice versa, and likewise for . If suppose first that . Then, using the fact that P is superharmonic and is harmonic in , as well as that both potentials have the same boundary values, we obtain from the maximum principle for harmonic functions that unless If we compare the boundary values of on Using the maximum principle for harmonic functions and (2) , we once again get a contradiction to (2) unless Interchanging the roles of , we obtain in all cases. The proof of Theorem 1 is now complete. □
Remark 2.
More sophisticated arguments in [15] yield that is in an open neighborhood of Fix and let vary in Then, from (19), the maximum principle for harmonic functions, and in Lemma 2, we see that
Using this fact, it is easily seen from Theorem 1 and the maximum principle for harmonic functions that M decreases as a function of so Moreover, as In view of our conjecture and results, it appears likely that
However, so far, we have not been able to prove this.
In , we can prove (38) without relying on [2], as follows: From (4) and (5), and a Schwarz–Christoffel-type argument, we have
where If and denotes the extremal potential corresponding to Then, (39) holds for this potential with b replaced by . Moreover,
Also, from (39), for we see that if then
Using (40) and (41), we can show that
Indeed, from the fact that is strictly decreasing on and the first derivative test for maxima, we find first that the maximum in (42) cannot occur at some Moreover, based on the same reasoning and (40) and (41), this maximum cannot occur when Finally, the integral in (42) for From (42) and (9), as well as Baernstein’s Theorem (mentioned after (9)), one can conclude that (3) in is valid for
In view of the above results, one wonders if in it is true that for
Remark 3.
Another question of interest to us is to what extent does Theorem 1 generalize to other PDEs? For example, can one replace the harmonic in Theorem 1 by the p-harmonic when . To be more specific, if does there exist a super p-harmonic function on with as satisfying of Theorem 1 for some choice of and (see [16] for relevant definitions)?
4. Proof of Proposition 1
Proposition 1.
whenever
Proof.
Since is symmetric in we assume as we may that Differentiating (44), we obtain
where
From(44) we see that (45) can be rewritten as
Taking second partials in (46), we have
Since h is symmetric in , we can interchange in (45) and (46) to get
Moreover, as in (46)
Adding in (47) to in (49), we get
Finally, we arrive at
From (47)–(51), we conclude that to finish the proof of Proposition 1, we need to show that the right-hand side of (51) is negative. To do this, we let and use the beta function to calculate in terms of our previous notation:
where is the Gamma function and when k is a positive integer with
In terms of this notation, and (47)–(54), we get at
if the term in the sum for D is
Let be the l th nonzero coefficient multiplying in If again then
Also
Using (57) and (58), we deduce that the part of the sum in (55) involving in D is
From (56) and (59), we get
Fix l as a positive integer and once again set As in the case , we see from (52)–(54) that
Also
Using (55) and (61), we see that the terms in D involving times are
We claim that
where Note from (60) that (64) is true when Proceeding by induction, assume (64) is true for some positive integer l. We need to show that
Now
From (66), we get for the term in (65)
Next, we observe from (66) for the term in (65) that
Hence, adding the term in (65) to the right-hand side of (68),
Finally, adding the right-hand sides of (67) and (69), we find from (65) and induction that claim (64) is true. From the definition of , we see that these functions have absolutely convergent series involving powers of and thereupon that (64) converges to Therefore, , and it follows from (55) that Proposition 1 is valid when is a positive integer. □
5. Proof of Theorem 2
Proof.
Recall from Section 1 that if and is the corresponding extremal potential satisfying of Theorem 1, then P is harmonic in so
and (3) holds whenever .
Thus, to prove Theorem 2, we show for d as in (70) and that
To do so, we first note for any that in as follows from the minimum principle for potentials and the fact that any two points on are at most distance two apart. Also, in [9], we showed the existence of in Theorem 2 whenever Thus, exists when d is as in (70).
The proof of (71) is by contradiction. For ease of writing, we put and write in spherical coordinates, which is permissible, since both functions are symmetric about the axis. Also, for fixed let be as defined in Theorem 1 relative to P while are defined relative to As in (14)–(17), and from Theorem 1, we deduce the existence of positive Borel measures, on with total mass corresponding to respectively. has its support in while has its support in Moreover,
We note from Theorem 2 that are continuous in the extended sense and non-increasing on whenever From this fact, (8) and (9), and d we deduce that to prove (71), it suffices to show
whenever . Moreover, from the Baernstein maximum principle (see the discussion after (8)), we need only prove (73) when To do this, observe that if (73) is false for some when then there exists with
Since is strictly decreasing on with on on and the integrals in (73) are equal when we see from the first derivative test in calculus that , and we may assume
To get a contradiction, we note from Proposition 1 and (72) that if
On the other hand, from Proposition 1, we have for
Using (72) and (77), we see for that
Combining (76) and (78), we get
since Thus, is increasing on so
since P is strictly decreasing on with Letting , we arrive at a contradiction to (74). From this contradiction, we conclude Theorem 2. □
Remark 4.
Conjecture 1 when or even an analogue of (10) when seems difficult. If the main problem is that the proposed extremal potential, must have mass at points on where it assumes both its maximum and minimum values on This splitting of the mass seems to rule out an immediate proof of Conjecture 1 using Baernstein’s ∗ function. A simpler problem in view of (9) is to show for
whenever Note that a positive answer to (81) would imply an analogue of (10) when and also (38).
To indicate our efforts in trying to prove (81), we first observe that it suffices to prove (81) when p is symmetric about the axis, so when Also, using the Baernstein ∗ function, as mentioned earlier, we need only prove (81) when Next, as in (15), we see that
where h is as in (16) and is a positive Borel measure on with Given it follows from the Fubini theorem and (82) that
where
We note that is continuous on (in fact, Hölder continuous), so from a theorem on weak convergence of measures, we deduce that
for some As for we can prove the following when
Lemma 3.
Let be the positive Borel measure on corresponding to in (85). Then,
Proof.
The proof of (86) is by contradiction and essentially given in Lemma 4 of [9]. To briefly outline the argument, if (86) is false, then using lower semi-continuity of one can assume there exists with on and For sufficiently small , and define by
where is to be chosen. One can verify from (87) that is an increasing function of on For let
Next, since is an increasing function of on , it follows that is a potential symmetric about the axis. A lengthy calculation then gives a sufficiently large B that and
From (89), we obtain
which contradicts (85), and so Lemma 1 is true. □
Remark 5.
Using (86) of Lemma 3, one can get as in [9] that for some
To use the above ‘mass moving’ argument further in proving (81) appears difficult since from the mean value property for harmonic functions,
Thus, moving mass inside (0, ), as above, would create points in this interval where and so perhaps not imply (90).
The lengthy calculation mentioned above was to show that
whenever and Originally in [9], after many months of trying, I had just proved (92) for n = 3, so (10) was just valid in Still, I submitted my paper to the Proc. of LMS and after a few months received a handwritten report from the referee (Walter Hayman) to the effect that I should try using the substitution,
to simplify my calculations. Using this observation, I was eventually able to prove (92) and after that use the above contradiction argument to get (10) in , n ≥ 3. Note that Proposition 1 is in the same spirit as (92).
Funding
This research received no external funding.
Data Availability Statement
No new data were created or analyzed in this study. Data sharing is not applicable to this article.
Conflicts of Interest
The author declares no conflicts of interest.
References
- Duren, P. Univalent Functions. In Grundlehren der Mathematischen Wissenschaften; Springer: New York, NY, USA, 2010; Volume 259. [Google Scholar]
- Barnard, R.; Lewis, J. Subordination Theorems for Some Classes of Starlike Functions. Pac. J. Math. 1975, 56, 333–366. [Google Scholar] [CrossRef][Green Version]
- Barnard, R.; Lewis, J. Coefficient Bounds for Some Classes of Starlike Functions. Pac. J. Math. 1975, 56, 325–331. [Google Scholar] [CrossRef][Green Version]
- Gariepy, R.; Lewis, J. A Maximum Principle with Applications to Subharmonic Functions in N-Space. Ark. Math. 1974, 12, 253–266. [Google Scholar] [CrossRef]
- Hayman, W.K. Subharmonic Functions, Volume 2. In London Mathematical Society Monographs; Academic Press: Cambridge, MA, USA, 1989; Volume 20. [Google Scholar]
- Baernstein, A., II; Drasin, D.; Laugeson, R. Symmetrization in Analysis; Cambridge University Press: Cambridge, UK, 2019. [Google Scholar]
- Baernstein, A., II. Proof of Edrei’s Spread Conjecture. Proc. Lond. Math. Soc. 1973, 26, 418–434. [Google Scholar] [CrossRef]
- Baernstein, A., II; Taylor, B.A. Spherical Rearrangements, Subharmonic Functions, and ∗ Functions in n Space. Duke Math. J. 1976, 43, 245–268. [Google Scholar] [CrossRef]
- Lewis, J. An n-Space Analogue of a Theorem of Suffridge. Proc. Lond. Math. Soc. 1975, 30, 444–464. [Google Scholar] [CrossRef]
- Suffridge, T. A Coefficient Problem for a Class of Univalent Functions. Mich. Math. J. 1962, 16, 33–42. [Google Scholar] [CrossRef]
- Hayman, W.K.; Kennedy, P.B. Subharmonic Functions, Volume 1. In London Mathematical Society Monographs; Academic Press: Cambridge, MA, USA, 1976; Volume 9. [Google Scholar]
- Evans, L.C. Partial Differential Equations. In Graduate Studies in Mathematics, 2nd ed.; American Mathematics Society: Providence, RI, USA, 2010; Volume 19. [Google Scholar]
- Kenig, C.; Jerison, D. Boundary Behavior of Harmonic Functions in Non-tangentially Accessible Domains. Adv. Math. 1982, 46, 80–147. [Google Scholar] [CrossRef]
- Silva, D.D.; Savin, O. C∞ Regularity of Certain Thin Free Boundaries. Indiana Univ. Math. J. 2015, 64, 1575–1608. [Google Scholar] [CrossRef]
- Silva, D.D.; Savin, O. Boundary Harnack Estimates in Slit Domains and Applications to Thin Free Boundary Problems. Rev. Mat. Ibero 2016, 32, 891–912. [Google Scholar] [CrossRef]
- Heinonen, J.; Kilpeläinen, T.; Martio, O. Nonlinear Potential Theory Degenerate Elliptic Equations; Dover Publications: Mineola, NY, USA, 2006. [Google Scholar]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content. |
© 2025 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access article distributed under the terms and conditions of the Creative Commons Attribution (CC BY) license (https://creativecommons.org/licenses/by/4.0/).