Skip to main content
Log in

Bounds on the growth of subharmonic frequently oscillating functions

  • Published:
Analysis and Mathematical Physics Aims and scope Submit manuscript

Abstract

We present a Phragmén–Lindelöf type theorem with a flavour of Nevanlinna’s theorem for subharmonic functions with frequent oscillations between zero and one. We use a technique inspired by a paper of Jones and Makarov.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8

Similar content being viewed by others

Notes

  1. Here, and anywhere else in this paper, we write \(A\lesssim B\) if there exists \(\alpha >0\) for which \(A\le \alpha \cdot B\). We write \(A\sim B\) if \(A\lesssim B\) and \(B\lesssim A\). If the constant depends on the dimension d, we will add a subscript d to each notation, i.e \(A\lesssim _d B\) and \(A\sim _d B\).

References

  1. Baranov, A., Hedenmalm, H.: Boundary properties of Green functions in the plane. Duke Math. J. 145, 1–24 (2008)

    Article  MathSciNet  Google Scholar 

  2. Bingham, N.H., Goldie, C.M., Teugels, J.L.: Regular Variation, Encyclopedia of Mathematics and its Applications. Cambridge University Press, Cambridge (1987)

    MATH  Google Scholar 

  3. Bishop, C., Peres, Y.: Fractals in Probability and Analysis, Cambridge Studies in Advanced Mathematics, vol. 162. Cambridge University Press, Cambridge (2017)

    Book  Google Scholar 

  4. Buhovsky, L., Glücksam, A., Logunov, A., Sodin, M.: Translation-invariant probability measures on entire functions. J. d’Anal. Math. 139, 1–33 (2019)

    Article  MathSciNet  Google Scholar 

  5. Carleson, L.: Selected Problems on Exceptional Sets. Van Nostrand, London (1967)

    MATH  Google Scholar 

  6. Hayman, W.K., Kennedy, P.B.: Subharmonic Functions Vol. 1 London Mathematical Society Monographs No. 9. Academic Press, London (1976)

    Google Scholar 

  7. Jones, P.W., Makarov, N.G.: Density properties of harmonic measure. Ann. Math. (2) 142(3), 427–455 (1995)

    Article  MathSciNet  Google Scholar 

Download references

Acknowledgements

The author is lexicographically grateful to Ilia Binder, Persi Diaconis, Eugenia Malinnikova, and Mikhail Sodin for several very helpful discussions. In particular, I would like to extend my gratitude to Chris Bishop who insisted on the connection between the problem presented here and the paper by Jones and Makarov, [7].

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Adi Glücksam.

Ethics declarations

Conflict of interest

The author declares that they have no conflict of interest.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supported in part by Schmidt Futures program.

Appendices

Appendices

1.1 Appendix A: A proof for Claim 1.2

The proof we present here is a concatenation of two inequalities. To state and prove these inequalities we will need the following definitions: Define the function

$$\begin{aligned} k_d(t):= {\left\{ \begin{array}{ll} \log \left( {t}\right) ,&{}\quad d=2\\ \frac{-1}{t^{d-2}},&{}\quad d\ge 3 \end{array}\right. }. \end{aligned}$$

Following [3], for every measure \(\nu \) we let

$$\begin{aligned} p_\nu (x):= & {} \int _{\mathbb R^d}^{}{k_d\left( {\left| {x-y}\right| }\right) }d{\nu }(y)\\ I(\nu ):= & {} \int _{\mathbb R^d}^{}{p_\nu (x)}d{\nu }(x)=\int _{\mathbb R^d}^{}{\int _{\mathbb R^d}^{}{k_d\left( {\left| {x-y}\right| }\right) }d{\nu }(y)}d{\nu }(x). \end{aligned}$$

The first is called the potential of the measure \(\nu \) and the second is called the energy of the measure \(\nu \). It is known that if the set E is compact, then there exists a probability measure \(\nu _0\) so that

$$\begin{aligned} I(\nu _0)=\underset{\nu }{\sup }\; I(\nu ) \end{aligned}$$

where the supremum is takes over all the Borel probability measures \(\nu \) which are supported on E. The measure \(\nu _0\) is called the equilibrium measure of E. For more information see, for example, Theorem 5.4 on page 209 of [6].

We are now ready to state these inequalities:

Claim 4.1

Let \(E\subset B\left( {0,\frac{1}{2}}\right) \) be a compact set with \(\lambda _{d-1}(E)>0\). If \(\nu _0\) is the equilibrium measure of E, then

$$\begin{aligned} \lambda _{d-1}(E)\lesssim _d\; \frac{-1}{I(\nu _0)}. \end{aligned}$$

The proof of this claim is a variation of the proof on p.94 in [3]. For the sake of completeness, and since the statement is not exactly the same, we bring here the full proof.

We will use the following result:

Lemma 4.2

(Frostman’s lemma) Let \(\varphi \) be a gauge function, i.e a positive, increasing function on \([0,\infty )\) with \(\varphi (0) = 0\). Let \(K\subset \mathbb R^d\) be a compact set with positive \(\varphi \)-Hausdorff content, \(\lambda _\varphi (K) > 0\). Then there is a positive Borel measure \(\mu \) on K satisfying that for every ball of radius r, B, \(\mu (B)\le C_d\cdot \varphi (r)\) while \(\mu (K)\ge \lambda _\varphi (K)\). The constant \(C_d\) depends on the dimension alone.

This lemma, and its proof can be found for example as Lemma 3.1.1, p.83 in [3], or as Theorem 1 on p.7 in [5].

Proof of Claim 4.1

Let \(\mu \) be the measure we obtain by using Lemma 4.2 with \(\varphi (t)=t^{d-1}\) and \(K=E\). We begin by bounding from bellow the potential of the measure \(\mu \). Then since \(-k_d\) is a monotone decreasing positive function on [0, 1] and \(diam(E)\le 1\)

$$\begin{aligned} -p_\mu (x):= & {} \int _{E}^{}{-k_d(\left| {x-y}\right| )}d{\mu }(y) \\\le & {} \underset{{n} ={0}}{\overset{\infty }{\sum }}\;\int _{\left| {x-y}\right|<2^{-n-1}}^{\left| {x-y}\right|<2^{-n}}{-k_d(\left| {x-y}\right| )}d{\mu }(y)\\\le & {} \underset{{n} ={0}}{\overset{\infty }{\sum }}\;\left\{ -k_d\left( {2^{-n-1}}\right) \left[ \mu \left( {\left\{ {\left| {x-y}\right|<2^{-n}}\right\} }\right) -\mu \left( {\left\{ {\left| {x-y}\right| <2^{-n-1}}\right\} }\right) \right] \right\} \\\le & {} C_d\underset{{n} ={0}}{\overset{\infty }{\sum }}\;2^{-n(d-1)} \left( {-k_d\left( {2^{-(n+1)}}\right) }\right) \\= & {} 2^{d-1}\cdot C_d\underset{{n} ={0}}{\overset{\infty }{\sum }}\;2^{-(n+1)(d-1)}\left( { -k_d\left( {2^{-(n+1)}}\right) }\right) \\\le & {} 2^{d-1}\cdot C_d\underset{{n} ={1}}{\overset{\infty }{\sum }}\;2^{-n(d-1)}\left( { -k_d\left( {2^{-n}}\right) }\right) \\= & {} {\left\{ \begin{array}{ll} 2^{d-1}\cdot C_d\underset{{n} ={1}}{\overset{\infty }{\sum }}\;\log \left( {2^n}\right) 2^{-n},&{} \quad d=2\\ 2^{d-1}\cdot C_d\underset{{n} ={1}}{\overset{\infty }{\sum }}\;\frac{2^{n(d-2)}}{2^{n(d-1)}},&{}\quad d\ge 3 \end{array}\right. } \end{aligned}$$

both series are convergent and bounded by 2. We conclude that \(-p_\mu (x)\le 2^dC_d\). Next, define the measure \(d\nu =\frac{d\mu }{\mu (E)}\). Then \(\nu \) is a probability measure supported on E, while

$$\begin{aligned} I(\nu )=\frac{1}{\mu (E)^2}I(\mu )=\frac{-1}{\mu (E)^2}\int _{\mathbb R^d}^{}{-p_\mu (x)}d{\mu }(x)\ge \frac{-2^dC_d}{\mu (E)}\ge \frac{-2^dC_d}{\lambda _{d-1}(E)}. \end{aligned}$$

By definition of equilibrium measure we see that

$$\begin{aligned} I(\nu _0)=\underset{\mu }{\sup }\; I(\mu )\ge I(\nu )\ge \frac{-2^dC_d}{\lambda _{d-1}(E)}\Rightarrow \lambda _{d-1}(E)\le \frac{-2^dC_d}{I(\nu _0)}, \end{aligned}$$

as \(I(\nu _0)<0\), concluding our proof. \(\square \)

Claim 4.3

Let \(E\subset B\left( {0,\frac{1}{2}}\right) \) be a compact set with \(\left| {I(\nu _0)}\right| <\infty \). Then

$$\begin{aligned} \omega (0,E;B(0,1){\setminus } E) > rsim _d\; \frac{-1}{I(\nu _0)}. \end{aligned}$$

Proof

Define the function \(w:B(0,1)\rightarrow \mathbb R\) by

$$\begin{aligned} w(x)= {\left\{ \begin{array}{ll} \omega \left( {x,E,B\left( {0,1}\right) {\setminus } E}\right) &{}\quad \xi \in B\left( {0,1}\right) {\setminus } E\\ 1 &{}\quad \text {otherwise} \end{array}\right. }. \end{aligned}$$

This function is super-harmonic, and by Poisson-Jensen formula for every \(x\in B\left( {0,1}\right) \), since \(w|_{\partial B(0,1)}\equiv 0\),

$$\begin{aligned} -w(x)= -\int _{B\left( {0,1}\right) }^{}{G_B\left( {y,x}\right) }d{\mu }_w(y), \end{aligned}$$
(⋆)

where \(G_B\) is Greens’s function on \(B=B\left( {0,1}\right) \), and \(\mu _w\) the Reisz measure of w, defined by \(d\mu _w=\frac{\Delta w}{\sigma (\partial B(0,1))}dm_d\), where \(\sigma _d\) is the surface measure in \(\mathbb R^d\). As \(E\subseteq B\left( {0,\frac{1}{2}}\right) \), there exists a constant \(c_d>0\) so that \(G_B(0,y)>c_d\) for every \(y\in E\), implying that

$$\begin{aligned} \omega \left( {0,E,B\left( {0,1}\right) {\setminus } E}\right)&=w\left( {0}\right) \overset{\text {by} \, (\star )}{=}\int _{E}^{}{G_B\left( {0,y}\right) }d{\mu }_w\left( {y}\right) \\&\ge c_d\cdot \int _{E}^{}{}d{\mu }_w\left( {y}\right) =c_d\cdot \left| \left| {\mu _w}\right| \right| _{}. \end{aligned}$$
(⋆⋆)

Next, let \(\nu \) denote the equilibrium measure of the set E, which is compact. For every measure \(\mu \) supported on E, by Frostman’s theorem:

$$\begin{aligned} \begin{aligned} \left| \left| {\mu }\right| \right| _{}&= \mu (E)=\frac{-1}{I(\nu )}\int _{\mathbb R^d}^{}{\int _{\mathbb R^d}^{}{-k_d(\left| {x-y}\right| )}d{\nu }(x)}d{\mu }(x)\\& > rsim _d \frac{-1}{I(\nu )}\int _{E}^{}{\int _{E}^{}{G_B\left( {x,y}\right) }d{\mu }(x)}d{\nu }(y), \end{aligned} \end{aligned}$$

for \(G_B\) Green’s function on \(B=B\left( {0,1}\right) \).

In particular, for \(\mu =\mu _w\) we get that

$$\begin{aligned} \begin{aligned} \left| \left| {\mu _w}\right| \right| _{}&\ge \frac{-1}{I(\nu )}\int _{E}^{}{\int _{E}^{}{G_B\left( {x,y}\right) }d{\mu }_w(y)}d{\nu }(x)\overset{\text {by} \, (\star )}{=}\frac{-1}{I(\nu )}\int _{E}^{}{w\left( {x}\right) }d{\nu }(x) \\&\overset{(a)}{=}\frac{-1}{I(\nu )}\int _{E}^{}{}d{\nu }(y)\overset{(b)}{=}\frac{-1}{I(\nu )}, \end{aligned} \end{aligned}$$

where (a) is since \(w|_E\equiv 1\), and (b) is since \(\nu \) is a probability measure supported on E.

Combining this with (⋆⋆), we see that

$$\begin{aligned} \omega \left( {0,E,B\left( {0,1}\right) {\setminus } E}\right) > rsim _d\;\left| \left| {\mu _w}\right| \right| _{}\ge \frac{-1}{I(\nu )}, \end{aligned}$$

concluding the proof. \(\square \)

Combining these two claims we see that

$$\begin{aligned} \omega (0,E; B(0,1){\setminus } E)\ge \frac{-C_1}{I(\nu )}\ge C_1\cdot C_2\cdot \lambda _{d-1}(E), \end{aligned}$$

concluding the proof of Claim 1.2.

1.2 Appendix B: Another proof of the main lemma

It is possible to prove the Main Lemma, Lemma 2.1, by using the Main Lemma in [7] with out repeating Jones and Makarov’s ingenious idea. This is done using the following lemma, suggested to the author by M.Sodin:

Lemma 4.4

Let u be a subharmonic function defined in a neighbourhood of \(Q_0:=\left[ - N, N\right] ^d\) for some \(N\gg 1\) and define \(Q=\left[ -\frac{N}{4\sqrt{d}},\frac{N}{4\sqrt{d}}\right] ^d\). Let \(E\subseteq Z_u \cap Q\) be a closed set. Then for every basic cube \(I\subset Q\) satisfying that \(\lambda _{d-1}(E\cap I)>\varepsilon _d>0\)

$$\begin{aligned} \frac{M_u(I)}{M_u(Q_0)}\lesssim \frac{\omega (I^*)}{\omega (E)}, \end{aligned}$$

where \(\omega (\cdot )=\omega (\infty ,\;\cdot \;;\mathbb R^d{\setminus } E)\) and \(I^*\) is a cube concentric with I having edge length which depends on the dimension alone.

In light of this lemma, and using the Main Lemma in [7], we get another proof for Lemma 2.1. For completeness, we bring here a reformulation of the Main Lemma in [7]:

Lemma 4.5

(The Main Lemma in [7]) Let \(E\subset B(0,1)\) be a compact set , and let \(\omega (\cdot )=\omega (\infty ,\cdot ;\mathbb R^d{\setminus } E)\). We subdivide \([0,1]^d\) into \(N^d\) cubes I with side length \(\frac{1}{N}\) and denote by \({\mathcal {G}}\) the whole collection of cubes. We define a subset \({\mathcal {E}}\subset {\mathcal {G}}\) (“empty” squares), as follows:

$$\begin{aligned} I\in {\mathcal {E}} \text { if }\lambda _{d-1}(E\cap I)\le \frac{\varepsilon _d}{N^{d-1}}, \end{aligned}$$

where \(\lambda _{d-1}\) is the \((d-1)\)-dimensional dyadic Hausdorff content. If \(\#{\mathcal {E}}\le c_d\cdot N^d\) for some absolute constant \(c_d\), then for at least half of the cubes \(I\in {\mathcal {G}}\) the following inequality holds:

$$\begin{aligned} \frac{\omega (I)}{\omega ([0,1]^d)}<\exp \left( {-\alpha _d\left( {\frac{N^d}{N+\#{\mathcal {E}}}\log ^d\left( {2+\frac{\#{\mathcal {E}}}{N}}\right) }\right) ^{\frac{1}{d-1}}}\right) , \end{aligned}$$

where \(\alpha _d\) is an absolute constant.

Alternative proof of Lemma 2.1

Following Lemma 4.4, it is enough to bound \(\omega (I^*)\) from above.

Let \(I\in {\mathcal {G}}{\setminus }{\mathcal {E}}\) be a basic cube and let \(I^*\) be the corresponding cube from Lemma 4.4. We rescale the set E by \(\frac{1}{\ell (I^*)N}\) so that rescaling \(I^*\), we get a cube of edge-length \(\frac{1}{N}\), and \(E\subset B(0,1)\). We may now apply the Main Lemma in [7], on the rescaled set, to conclude the proof. \(\square \)

1.3 The proof of Lemma 4.4

Let g denote Green’s function for the domain \(\Omega :=\mathbb R^d{\setminus } E\) with pole at \(\infty \). Since \(E\subseteq Q\subset B\left( {0,\frac{N}{2}}\right) \) and Green’s functions satisfy the subordination principle, for every \(\left| {x}\right| \ge N\)

$$\begin{aligned} g(x)\ge g_{\mathbb R^d{\setminus } B\left( {0,\frac{N}{2}}\right) }(\infty ,x)\ge \underset{\left| {x}\right| {=}N}{\min }\; g_{\mathbb R^d{\setminus } B\left( {0,\frac{N}{2}}\right) }(\infty ,x){=}\underset{\left| {x}\right| =2}{\min }\; g_{\mathbb R^d{\setminus } B(0,1)}(\infty ,x):= c_1, \end{aligned}$$

where \(c_1\) is a constant which depends on the dimension, but independent of N. By the maximum principle, for every subharmonic function, u, with \(E\subset Z_u\),

$$\begin{aligned} \frac{u(x)}{M_u(Q_0)}\le 1\le \frac{g(x)}{c_1}, \end{aligned}$$

implying that

$$\begin{aligned} \frac{M_u(I)}{M_u(Q_0)}\lesssim M_g(I). \end{aligned}$$

To conclude the proof we shall bound \(M_g(I)\) from above.

Let h be a subharmonic function and let I be a basic cube satisfying property (P2), and let \(x_I\in \partial I\) be a point satisfying that \(M_h(I)=h(x_I)\). By taking a square double in edge length, we may assume without loss of generality that \(dist(x_I,E\cap I)>\frac{1}{3}\). We note that following Nevanlinna’s first fundamental theorem (see for example [6, Theorem 3.19]), for every \(r>\sqrt{d}\)

$$\begin{aligned} M_h\left( {B\left( {x_I,\sqrt{d}}\right) }\right) \le \frac{r^{d-2}(r+\sqrt{d})}{(r-\sqrt{d})^{d-1}}\left( {h(x_I)+\underset{y\in E\cap I}{\max }G_{B(x_I,r)}(x_I,y)\cdot \mu _h(B(x_I,r))}\right) , \end{aligned}$$

and \(\mu _h\) is Riesz measure, defined by \(d\mu _h(x)=\frac{\Delta h(x)}{\sigma _d(\partial B(0,1))}dm_d(x)\), for \(\sigma _d\) the surface area measure in \(\mathbb R^d\).

If \(r=\alpha \cdot \sqrt{d}\), then for every d fixed

$$\begin{aligned}&\frac{r^{d-2}(r+\sqrt{d})}{(r-\sqrt{d})^{d-1}} =\frac{\alpha ^{d-2}d^{\frac{d-2}{2}}\cdot \sqrt{d}(\alpha +1)}{d^{\frac{d-1}{2}}(\alpha -1)^{d-1}}=\frac{\alpha ^{d-2}(\alpha +1)}{(\alpha -1)^{d-1}}=\frac{1+\frac{1}{\alpha }}{\left( {1-\frac{1}{\alpha }}\right) ^{d-1}}\searrow 1, \\&\quad \text { as } \alpha \rightarrow \infty . \end{aligned}$$

We conclude that for every \(\delta >0\) there exists \(\alpha \) so that

$$\begin{aligned} M_h\left( {B\left( {x_I,\sqrt{d}}\right) }\right) \le \left( {1+\delta }\right) \left( {h(x_I)+\underset{y\in E\cap I}{\max }G_{B(x_I,\alpha \sqrt{d})}(x_I,y)\cdot \mu _h(B(x_I,\alpha \sqrt{d}))}\right) . \end{aligned}$$
(⋆)

Next, since the basic cube I satisfies that \(\lambda _{d-1}(E\cap I)>\varepsilon _d>0\), following Observation 1.1, there exists \(\delta _d>0\) for which

$$\begin{aligned} h(x_I)<M_h\left( {B\left( {x_I,\sqrt{d}}\right) }\right) (1-\delta _d). \end{aligned}$$

Combining this with (), we note that if we choose \(\alpha >0\) large enough (depending on the dimension d, and, in particular, on \(\delta _d\))

$$\begin{aligned}&M_h\left( {B\left( {x_I,\sqrt{d}}\right) }\right) \\&\quad \le \left( {1+\frac{\delta _d}{2}}\right) \left( {h(x_I)+\underset{y\in E\cap I}{\max }G_{B(x_I,\alpha \sqrt{d})}(x_I,y)\cdot \mu _h(B(x_I,\alpha \cdot \sqrt{d}))}\right) \\&\quad \le \left( {1+\frac{\delta _d}{2}}\right) \left( {(1-\delta _d)M_h\left( {B\left( {x_I,\sqrt{d}}\right) }\right) +c_d\cdot \mu _h(B(x_I,\alpha \cdot \sqrt{d}))}\right) , \end{aligned}$$

for some constant \(c_d\), which depends on the dimension alone.

This implies that if \(I^*\) is a cube centred at \(x_I\) with edge length \(2\alpha \sqrt{d}\) then

$$\begin{aligned} M_h(I)\le M_h\left( {B\left( {x_I,\sqrt{d}}\right) }\right) \le \frac{4c_d}{\delta _d}\cdot \mu _h(B(x_I,\alpha \cdot \sqrt{d}))\le \frac{4c_d}{\delta _d}\cdot \mu _h(I^*). \end{aligned}$$

In particular, for \(h=g\) we see that

$$\begin{aligned} M_g(I)\lesssim \mu _g(I^*). \end{aligned}$$

To conclude the proof, we use the fact that for every cube J:

$$\begin{aligned} \mu _g(J)=\sigma _d\left( {B(0,1)}\right) \cdot \omega (\infty ,J;\Omega ), \end{aligned}$$

for \(\sigma _d\) the surface area in \(\mathbb R^d\), implying that

$$\begin{aligned} \frac{M_u(I)}{M_u(Q_0)}\lesssim M_g(I)\lesssim \mu _g(I^*)\sim _d\omega (\infty ,I^*,\mathbb R^d\setminus E)\le \frac{\omega (I^*)}{\omega (E)}. \end{aligned}$$

\(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Glücksam, A. Bounds on the growth of subharmonic frequently oscillating functions. Anal.Math.Phys. 11, 69 (2021). https://doi.org/10.1007/s13324-021-00489-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s13324-021-00489-1

Navigation