Saldana
Saldana
doi:10.3934/dcds.2024084
Vı́ctor Hernández-Santamarı́a 1 ,
∗1
Luis Fernando López Rı́os 2 and Alberto Saldaña
1 Instituto de Matemáticas, Universidad Nacional Autónoma de México,
1
2 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
naturally appears. We also mention the following very recent works: the Cauchy
problem with L∆ is studied in [9]; higher-order expansions of the fractional Lapla-
cian are considered in [7]; a characterization of the logarithmic Laplacian via a
local extension problem on the (N + 1)-dimensional upper half-space in the spirit
of the Caffarelli-Silvestre extension is obtained in [8]; and a finite element method
is analyzed and implemented in [23] to approximate solutions of (1).
The kernel in (2) is sometimes called of zero-order, because it is a limiting case for
hypersingular integrals. As a consequence, the regularizing properties of this type of
operators are very weak and are a subject of current research. Interior regularity of
(bounded weak) solutions to (1) has been studied in [6,10,21,27]; in particular, it is
known that if f ∈ L∞ (Ω) then u ∈ C(Ω). Furthermore, the modulus of continuity of
u can be characterized: u belongs to the space of α-log-Hölder continuous functions
in RN for some α ∈ (0, 1) (see the definition of Lα (RN ) in (20) below); in particular,
|u(x) − u(y)|
kukLα (RN ) = kukL∞ (RN ) + sup < ∞, (3)
x,y∈RN `α (|x − y|)
x6=y
We refer to [23, Figure 1] for the numerical approximation of the torsion function
in different intervals.
The proofs of Theorems 1.1 and 1.2 are done by constructing suitable barriers
and by using a comparison principle in small domains. However, in the case of the
logarithmic Laplacian, this is a highly nontrivial task, because there is no easy choice
for a suitable barrier; in particular, a closed formula for the torsion function in any
ball is not available and one cannot argue as in [30] (this is also the reason why (4)
is only established for τ ∈ (0, 12 ) in [10]). To overcome this difficulty, we use the
following strategy. First, we construct a barrier in an open interval I = (0, 2) that
1
behaves as ` 2 (x) for x close to 0 (see (26)). Then, via sharp direct computations, we
show that its logarithmic Laplacian is bounded in I and that it is positive close to
0, see Theorem 2.4. Afterwards, we use this barrier to construct higher-dimensional
barriers in half balls (see Figure 1 and (44)). Using the calculations in the one-
dimensional case and a Leibniz-type formula (see Section A.2), we show that this
new function also has a bounded logarithmic Laplacian that is positive close to the
origin (see Theorem 2.19).
Finally, to obtain a suitable barrier to characterize domains satisfying the (uni-
form) exterior sphere condition, we use the following Kelvin transform formula,
which is of independent interest. Let x, x0 ∈ RN and R > 0, the inversion of x with
4 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
1.4
1.2
1.0
0.8
0.6
0.4
0.2
Note that formula (10) is different from the one obtained for the fractional Lapla-
cian (see, for example, [3] or [1, Proposition 2]). For the logarithmic Laplacian, the
geometry of the domain has a stronger influence in the formula, and this is repre-
sented in the last two summands in (10).
Using the initial barrier and the Kelvin transform, we obtain a new function
which has a bounded logarithmic Laplacian and which has the optimal regularity
at the curved part of the boundary (Proposition 3.4), see Figure 2.
1.0
0.5
-0.5
-1.0
With this barrier one has the main ingredient to characterize the optimal bound-
ary regularity. The other important ingredient is the comparison principle. Al-
though the operator L∆ does not satisfy the maximum principle in general, in [10,
Corollary 1.9] (see Theorem 4.1 in Section 4) it is shown that it holds in sufficiently
small domains. With this and with suitable scalings (see Section A.1) one can adapt
the method of barrier functions.
As a byproduct of our approach, we can show the following Hopf-type lemma for
the logarithmic Laplacian (following the ideas in [16]). Let EL be the bilinear form
6 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
q < ∞ as s → 0+ , where v is a positive solution of (13). See [2] for more details.
Note that these problems are sublinear. In the superlinear case (ps > 2) uniqueness
is known to be false in general (see [12–15, 17, 18] for uniqueness and multiplicity
results for positive solutions of fractional problems). Equation (13) with µ < 0
(which would be the superlinear case in the logarithmic setting) has been studied
in [24], but nothing is known yet about the uniqueness or multiplicity of positive
solutions and our techniques cannot be used in this setting.
To close this introduction, we mention that we believe a similar strategy can also
be used to study the sharp boundary behavior of more general integral operators
associated with zero-order kernels, as those considered in [6,11, 19,21], for instance.
The paper is organized as follows. In Section 2 we construct the initial barriers
described in Figure 1. Section 3 contains the proof of Proposition 1.3 regarding the
Kelvin transform and the construction of the barrier portrayed in Figure 2. Theo-
rem 1.1 is shown in Section 4, whereas the proofs of Theorem 1.2 (optimal regularity
OPTIMAL BOUNDARY REGULARITY FOR THE LOGARITHMIC LAPLACIAN 7
for the torsion problem), Theorem 1.4 (Hopf-type lemma for the logarithmic Lapla-
cian), and the uniqueness of positive solutions of (13) (Theorem 5.5) are contained
in Section 5. Finally, we include an Appendix with some useful results regarding
scalings and a Leibniz-type formula for the logarithmic Laplacian of a product.
Notation. We use Br (x) to denote the open ball of radius r > 0 centered at
x ∈ RN . If x = 0, then we simply write Br := Br (0). We also set Sr (x) = ∂Br (x),
Sr := Sr (0), and
N
2π 2
σN := |S1 | =
Γ( N2 )
to denote the surface measure of the unit sphere. For U ⊂ RN , we use U c := RN \U.
The constants involved in the definition of L∆ are given by
N
cN := π − 2 Γ( N2 ), ρN := 2 ln 2 + Ψ( N2 ) − γ, and γ := −Γ0 (1). (14)
0
Here γ is also known as the Euler-Mascheroni constant and Ψ := ΓΓ is the digamma
function.
Next, following [10], we introduce the variational framework for L∆ . Let Ω ⊂ RN
be an open bounded set and let H(Ω) be the Hilbert space given by
(
|u(x) − u(y)|2
Z Z
H(Ω) := u ∈ L2 (RN ) : dx dy < ∞
x,y∈RN
|x − y|N
|x−y|≤1
)
and u = 0 in RN \ Ω (15)
`(r)
1
10
| r
For α > 0, we also define the α-log-Hölder Banach space (see [6, Lemma 7.1]) by
Lα (Ω) := {u : Ω → R : kukLα (Ω) < ∞}, (20)
where
|u(x) − u(y)|
kukLα (Ω) := kukL∞ (Ω) + [u]Lα (Ω) , [u]Lα (Ω) := sup α
,
x,y∈Ω ` (|x − y|)
x6=y
∞
and k · kL∞ (Ω) is the usual norm in L (Ω). We also set kuk∞ := kukL∞ (RN ) .
We shall use the following semi-homogeneity property for the modulus of conti-
nuity `.
Lemma 1.5 (Lemma 3.2 in [6]). There is c > 0 such that
`(λr) ≥ c `(λ)`(r) for all r, λ > 0.
OPTIMAL BOUNDARY REGULARITY FOR THE LOGARITHMIC LAPLACIAN 9
2. An initial barrier. The main goal of this section is to show the following result.
Let N ≥ 1 and let
B2+ := B2 ∩ RN
+, where RN
+ := {x ∈ R
N
: x1 > 0}. (21)
1 +
Theorem 2.1. There is a function w ∈ L 2 (RN ) ∩ C ∞ (RN \RN
+ ) ∩ H(B2 ) such that
N +
w = 0 in R \B2 and
L∆ w ∈ L∞ (B2+ ).
Moreover, there is c > 0 and δ > 0 such that
1 1
c ` 2 (x1 ) < w(x) < c−1 ` 2 (x1 ) and L∆ w(x) > c for all x ∈ B2+ ∩ Bδ .
To show Theorem 2.1 we construct explicitly the function w and estimate its
logarithmic Laplacian with direct calculations.
We show first some auxiliary lemmas. Let Ω ⊂ RN be an open bounded set.
Lemma 2.2. Let V ∈ L∞ (RN ) be such that V = 0 in RN \Ω and
V (z) − V (z + y)
Z
sup dy < C (22)
z∈Ω B1 |y|N
for some C > 0. Then L∆ V ∈ L∞ (Ω).
Proof. Let x ∈ Ω, then
V (x) − V (x + y)
Z Z
V (y)
|L∆ V (x)| = cN N
dy − cN dy + ρN V (x)
B1 |y| Ω\B1 (x) |y − x|N
≤ cN C + kV k∞ (cN |Ω| + |ρN |) < ∞.
Now, we split the proof of Theorem 2.1 in two cases: the (simpler) one-dimensional
case and the higher-dimensional case.
10 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
because u(0) = 0. Using that t 7→ |u0 (t)| is decreasing in (0, 41 ), it follows that
ε 1
< ε((1 − s)y + 1) < 2ε < for s ∈ (0, 1) and y ∈ (− 12 , 1),
2 4
and then
1 1 1
|u(ε) − u((y + 1)ε)|
Z Z Z
dy ≤ ε |u0 (ε((1 − s)y + 1))| ds dy
− 12 |y| − 21 0
3
≤ ε|u0 ( 14 ε)| = o(1)
2
as ε → 0.
Lemma 2.7. It holds that
Z 1 !
1 1 1
I := I(ε) := p −p dy = o(1)
ε y − ln (y + ε) + ln 2 − ln(y) + ln 2
as ε → 0.
Proof. Using the change of variables z = − ln y (y = e−z , dy = −e−z dz) we obtain
that
Z − ln ε
1 1
I= p −√ dz = o(1)
−z
− ln(e + ε) + ln 2 z + ln 2
0
as ε → 0. Indeed, consider
!
1 1
Fε (z) := p −√ χ{z<− ln ε} (z),
− ln(e−z + ε) + ln 2 z + ln 2
for z ∈ (0, ∞) and ε ∈ (0, 1). Observe that
1 1 1 1
√ − √ ≤ 3/2 (t − s), ln t − ln s ≤ (t − s) for all 0 < s ≤ t,
s t s s
by the mean value theorem. By applying these estimations to Fε we deduce that
−z −z
z + ln(e + ε) ln(e + ε) − ln (e−z )
Fε (z) ≤ 3/2
= 3/2
2 [− ln(e−z + ε) + ln 2] 2 [− ln(e−z + ε) + ln 2]
ez ε 1
≤ 3/2
≤ 3/2
for z ≤ − ln ε.
2 [− ln(e−z + ε) + ln 2] 2 [− ln(e−z + ε) + ln 2]
(27)
So let us define, for any ε ≥ 0,
1
Gε (z) := 3/2
, z ∈ R.
2 [− ln(e−z + ε) + ln 2]
Observe that Gε (z) &R ∞G0 (z) as εR → 0 for z ∈ R, so the monotone convergence
∞
theorem implies that 0 Gε (z) & 0 G0 (z) as ε → 0. On the other hand, by (27),
Fε (z) ≤ Gε (z) for z ∈ R and Fε (z) → F0 (z) = 0 as ε → 0 for zR ∈ R, so the
∞
generalized Lebesgue dominated convergence theorem implies that 0 Fε → 0 as
ε → 0, as claimed.
Lemma 2.8. It holds that
Z 1+ε
u(y) √ √
J2 = J2 (ε) := − dy = 2 ln 2 − 2 − ln ε + o(1) as ε → 0. (28)
2ε |ε − y|
12 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
Theorem 2.9. The function u given by (26) belongs to C ∞ ((0, ∞)) and
Z ε+1
u(ε) − u(y) √
J = J(ε) := dy = 2 ln 2 + o(1) as ε → 0+ . (31)
ε−1 |y − ε|
Proof. That u ∈ C ∞ ((0, ∞)) is clear. Moreover, since u = 0 in (−∞, 0),
Z ε+1 Z 0
u(ε) − u(y)
J= dy + u(ε) |ε − y|−1 dy
0 |ε − y| ε−1
Z ε+1
u(ε) − u(y)
= dy + u(ε)(− ln ε) (32)
0 |ε − y|
for ε > 0 small. Let J1 and J2 as in Lemmas 2.6 and 2.8. Then,
Z 1+ε Z 2ε Z 1+ε
u(ε) − u(y) u(ε) − u(y) u(ε) − u(y)
dy = dy + dy
0 |ε − y| 0 |ε − y| 2ε |ε − y|
Z 1+ε
1
= J1 + J2 + u(ε) dy = J1 + J2 + u(ε)(− ln ε).
2ε |ε − y|
(33)
Now the claim (31) follows by (32), (33), Lemma 2.6, Lemma 2.8, and the fact that
√
2u(ε)(− ln(ε)) = 2 − ln ε + o(1) as ε → 0+ .
Z Z 1
≤ kuk∞ ku0 kL∞ (U ) + khkL∞ ((−1,1)) dy dx < ∞,
U −1
ζ 1 ζ
(u(y + x) − u(x))u(y + x)
Z Z Z
A2 := dy dx ≤ 2kuk2∞ − ln(x) dx < ∞.
0 x |y| 0
Moreover, using that u = 0 in (−∞, 0) and Lemma 2.6,
Z ζZ x
(u(y + x) − u(x))u(y + x)
A3 := dy dx
0 −1 |y|
Z ζZ x
|u(x) − u(y + x)|
≤ kuk∞ dy dx < ∞.
0 −x |y|
Furthermore, using that u = 0 in (−∞, 0),
Z 0 Z 1
(u(y + x) − u(x))u(y + x)
A4 := dy dx
−ζ −1 |y|
Z 0 Z 1
(u(y + x) − u(x))u(y + x)
= dy dx
−ζ −x |y|
Z 0
≤ 2kuk2∞ − ln(−x) dx < ∞.
−ζ
Then,
1
(u(y + x) − u(x))u(y + x)
Z Z
A := dy dx = A1 + A2 + A3 + A4 < ∞. (34)
R −1 |y|
Recall the definition of the norm k · k given in (16). Then,
Z Z x+1
2 (u(x) − u(y))(u(x) − u(y))
2kuk = dy dx
R x−1 |x − y|
Z Z x+1 Z Z x+1
(u(x) − u(y))u(x) (u(y) − u(x))u(y)
= dy dx + dy dx.
R x−1 |x − y| R x−1 |x − y|
Using Fubini’s theorem, a change of variables (z = x − y), and (34),
Z Z x+1 Z Z y+1
(u(x) − u(y))u(x) (u(x) − u(y))u(x)
dy dx = dx dy
R x−1 |x − y| R y−1 |x − y|
Z Z 1
(u(z + y) − u(y))u(z + y)
= dz dy = A < ∞.
R −1 |z|
Similarly,
Z Z x+1 Z Z 1
(u(y) − u(x))u(y) (u(z + x) − u(x))u(z + x)
dy dx = dz dx
R x−1 |x − y| R −1 |z|
= A < ∞.
Thus kuk < ∞. Since we also know that u ∈ L2 (R), it follows that u ∈ H((0, 2)) as
claimed.
We are ready to show Theorem 2.4.
Proof of Theorem 2.4. By Lemma 2.10 and by the definition of u it holds that u ∈
1
H((0, 2))∩C ∞ ((0, ∞))∩L 2 (RN ). By Theorem 2.9 and the fact that u ∈ C ∞ ((0, ∞))
we have that (22) holds (with Ω = (0, 2) and V = u). Then, by Lemma 2.2, we
14 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
√
have that L∆ u ∈ L∞ ((0, 2)). Finally, by
√ Lemma 2.3 (with σ = ln 2), Lemma 2.5,
and Theorem 2.9 we have that L∆ u ≥ ln 2 in (0, δ) for some δ > 0.
2.2. The higher-dimensional case. In this section we extend the ideas of the
one-dimensional case to higher dimensions. Let N ≥ 2,
RN
+ := {x ∈ R
N
: x1 > 0},
and let ζ be such that
√ r ! N1
1 N ln 2 4
ζ< 1+ ln − 1 . (35)
2 4 3
1
Note that, since (1 + N x) N < 1 + x for x ≥ 0, inequality (35) implies (24); namely,
√ r
ln 2 4 1
0<ζ< ln < .
4 3 4
Let u and ϕ as in (25) and (26); namely, ϕ ∈ Cc∞ (R) is an even function such
that
ϕ = 1 in (−1 − ζ, 1 + ζ), ϕ = 0 in R\(−1 − 2ζ, 1 + 2ζ), 0 ≤ ϕ ≤ 1 in R
− 12
and u(x) := (− ln x2 ) ϕ(x)χ[0,∞) (x) for x ∈ R. Let
ϕ(x1 )
V ∈ L1loc (RN ) be given by V (x) := u(x1 ) = p χ[0,∞) (x1 ), (36)
− ln x21
see Figure 1 (right). For ε ∈ (0, 14 ), let
xε = (ε, 0, . . . , 0) ∈ RN and Uε := {y ∈ B1 (0) : y1 ∈ (−ε, ε)}.
For N ≥ 1, we use σN to denote the surface measure of the sphere in RN , namely,
N
σN = Γ(NNπ+1)
2
.
2
We show first some auxiliary lemmas.
Lemma 2.11.
V (xε ) − V (y + xε )
Z
J1 = J1 (ε) := dy = o(1) as ε → 0.
Uε |y|N
Proof. Let B := {y 0 ∈ RN −1 : |y 0 | < 1}. Using that u is increasing in (0, ε) for
ε > 0 small and that u(0) = 0,
Z Z − 2ε − N2
ε ε2
|u(ε) − u(y1 + ε)|
Z
0 0 2
J1,1 := N dy1 dy ≤ u(ε) + |y | dy 0
B −ε (y12 + |y 0 |2 ) 2 B 2 4
Z 1 2 − N2
ε ε
= u(ε)σN −1 + r2 rN −2 dr
2 0 4
Z ∞
− N
≤ u(ε)σN −1 1 + r2 2 rN −2 dr = o(1)
0
Z Z ε Z 1
y1
≤ |u0 (ty1 + ε)| N dt dy1 dy 0
B − 2ε 0 (y12 + |y 0 |2 ) 2
Z Z ε
y1
≤ |u0 ( 2ε )| N dy1 dy 0
B − 2ε (y12 + |y 0 |2 ) 2
Z 1Z ε
y1
= σN −1 |u0 ( 2ε )| N rN −2 dy1 dr
0 − 2ε (y12 + r2 ) 2
∞ 1
y1 rN −2
Z Z
≤ σN −1 ε|u0 ( 2ε )| N dy1 dr
0 − 12 (y12 + r2 ) 2
∞
rN −2
Z
3
= σN −1 ε|u0 ( 2ε )| N dr = o(1)
2 0 (1 + r2 ) 2
as ε → 0+ , where we used that ε|u0 ( 2ε )| → 0 as ε → 0+ . Then,
|V (xε ) − V (y + xε )|
Z
dy ≤ J1,1 + J1,2 = o(1) as ε → 0 (37)
Uε |y|N
and this ends the proof.
h i
For ε ∈ (0, 14 ) and η ∈ √1 , 1
2
, let
Z ∞ Z ln t−ln ε
2 −N 1
= σN −1 |t + 1| 2 √dz dt − R(η)
ε
η ln t−ln η ln 2 − ln t + z
Z ∞ Z − ln ε !
2 −N 1
= σN −1 |t + 1| 2 √ dz dt − R(η)
ε
η − ln η ln 2 + z
√ p Z ∞ N
= 2σN −1 ln 2 − ln ε − ln 2 − ln η |t2 + 1|− 2 dt − R(η)
η −1 ε
√ !
√ πΓ N 2−1
p
= ln 2 − ln ε − ln 2 − ln η σN −1 + o(1) − R(η)
Γ N2
√ p
= σN − ln ε − σN ln 2 − ln η − R(η) + o(1),
ηez
εez
-1
p
− ln η z
z ln t − ln ε
ln t − ln η
− ln η
ε
η
p p t
1
because
√ N −1
N N
πΓ π2 π2
σN −1 N
2 =2 N =N N = σN .
Γ 2 Γ( 2 ) Γ( 2 + 1)
OPTIMAL BOUNDARY REGULARITY FOR THE LOGARITHMIC LAPLACIAN 17
as ε → 0, whereas
Z
V (y + xε ) N
0< dy ≤ 2 2 kV k∞ |B1 |. (42)
B1 \(Uε ∪Qε, √1 ) |y|N
2
18 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
where ϕ is given in (25). Recall that B2+ is the half ball of radius 2 given in (21).
V (z) − V (z + y)
Z
sup dy < C
z∈B + B1 |y|N
2
for some C > 0, as a simple argument by contradiction shows. From this estimate
and Lemma 2.2 we deduce that L∆ V ∈ L∞ (B2+ ). On the other hand, by the
Leibniz-type formula (78),
Furthermore, recall that V has support on the strip {|y1 | < 1 + 2ζ} (see (36)). Since
x1 ∈ (0, 2), we have that {|x1 + y1 | < 1 + 2ζ} ⊂ {|y1 | < 4} and then
Z
V (x + y)
A3 = cN ϕ(x) dy
RN \B1 |y|N
Z
≤ cN kV k∞ kϕk∞ |y|−N dy
(RN \B1 )∩{|x1 +y1 |<1+2ζ}
Z
≤ cN kV k∞ kϕk∞ |y|−N dy
{|y 0 |<1}∩{|y1 |<4}∩{|y|>1}
Z Z !
0 −N 0
+ |y | dy dy1
{|y1 |<4} {|y 0 |>1}
Z ∞
≤ cN kV k∞ kϕk∞ {|y 0 | < 1} ∩ {|y1 | < 4} + {|y1 | < 4} σN −1 ρ−2 dρ
1
< ∞,
where y = (y1 , y 0 ) ∈ R × RN −1 .
Lemma 2.17. It holds that
σN √
Z
Ve (y)
N
dy < ln 2 for ε ∈ (0, ζ).
RN \B1 (xε ) |y − xε | 4
Proof. By (35),
ϕ(y1 )χ(0,∞) (y1 )
Z Z
Ve (y) 1
N
dy = y1 dy
|y − xε |N
p
RN \B1 (xε ) |y − xε | B1+2ζ \B1 (xε ) − ln 2
Z
1 1 1
≤q dy ≤ q |B1+2ζ \B1 (xε )|
1+2ζ B1+2ζ \B1 (xε ) |y − xε |N
− ln 2 − ln 1+2ζ 2
1 1 σN σN √
=q |B1+2ζ \B1 | ≤ q (1 + 2ζ)N − 1 < ln 2
− ln 1+2ζ 1+ 12 N 4
2 − ln 2
(Ve (y + x) − Ve (x))Ve (y + x)
Z Z
A2 := dy dx
K∩{x1 >0} B1 ∩{y1 >x1 } |y|N
20 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
Z Z 1 Z 1
N
≤ 2σN −1 kVe k2∞ |y12 + ρ2 |− 2 ρN −2 dy1 dρ dx
K∩{x1 >0} 0 x1
1 ∞
ρN −2
Z Z Z
≤ 2σN −1 kVe k2∞ y1−1 N dρ dy1 dx
K∩{x1 >0} x1 0 |1 + ρ2 | 2
√ N −1
Z
πΓ
= 2σN −1 kVe k2∞ N
2
− ln(x1 ) dx < ∞. (45)
2Γ 2 K∩{x1 >0}
Moreover,
(Ve (y + x) − Ve (x))Ve (y + x)
Z Z
A3 := dy dx
K∩{x1 >0} B1 ∩{y1 <x1 } |y|N
|Ve (x) − Ve (y + x)|
Z Z
≤ kVe k∞ dy dx < ∞,
K∩{x1 >0} B1 ∩{−x1 <y1 <x1 } |y|N
where the finiteness of A3 follows from (37) (note that (37) is stated for V instead
of Ve , but since Ve is given by (44), the bound easily extends to Ve ).
Furthermore, since Ve (x + y) = 0 if x1 + y1 < 0, arguing as in (45),
(Ve (y + x) − Ve (x))Ve (y + x)
Z Z
A4 := dy dx
K∩{x1 <0} B1 |y|N
(Ve (y + x) − Ve (x))Ve (y + x)
Z Z
= dy dx
K∩{x1 <0} B1 ∩{−x1 <y1 <1} |y|N
Z 1Z 1
ρN −2
Z
≤ 2σN −1 kVe k2∞ dx dρ dy1 dx
2 2 N
K∩{x1 >0} x1 0 |y1 + ρ | 2
Z 1 Z ∞
rN −2
Z
2 −1
≤ 2σN −1 kV k∞
e dx y1 dy1 dr
K∩{x1 >0} x1 0 |1 + r2 |N/2
Z
= CkVe k2∞ − ln(x1 )dx < ∞,
K∩{x1 >0}
(Ve (z + y) − Ve (y))Ve (z + y)
Z Z
= dz dy = A < ∞.
RN B1 |z|N
Similarly,
(Ve (y) − Ve (x))Ve (y)
Z Z
χB1 (x) (y) dy dx
RN RN |x − y|N
(Ve (z + x) − Ve (x))Ve (z + x)
Z Z
= dz dx = A < ∞.
RN B1 |z|N
Then kVe k2 < ∞. Since we also know that Ve ∈ L2 (RN ), we obtain that Ve ∈ H(B2+ )
as claimed.
The following result implies Theorem 2.1 in the case N ≥ 2.
Theorem 2.19. The function Ve given by (36) belongs to H(B2+ ) ∩ C ∞ (RN +) ∩
1
L 2 (RN ) and there is δ ∈ (0, 2) such that
σN √
L∆ Ve ∈ L∞ (B2+ ) and L∆ Ve (x) > ln 2 for x ∈ Bδ ∩ RN
+.
2
Proof. By Lemma 2.16 we know that L∆ Ve ∈ L∞ (B2+ ). By Lemma 2.18 and by
1
construction we have that Ve ∈ H(B2+ ) ∩ C ∞ (RN N
+ ) ∩ L (R ). Finally, by Theo-
2
rem 2.15,
Z
V (xε ) − V (y) √
J := N
dy > σN ln 2 + o(1) as ε → 0.
B1 (xε ) |y − xε |
Since, by (25), it holds that ϕ(|x|) = 1 for x ∈ B1+ζ , we have that Ve = V in B1+ζ
(with V given in (36)) and therefore
Ve (x) − Ve (y) V (x) − V (y)
Z Z
N
dy = dy for all x ∈ Bζ .
B1 (x) |y − x| B1 (x) |y − x|N
In particular,
Z
Ve (yε ) − Ve (y) √
N
dy = J > σN ln 2 + o(1) as ε → 0 (47)
B1 (yε ) |y − yε |
for all yε = (ε, y 0 ) where y 0 ∈ R√
N −1
is such that yε ∈ Bζ . Then, by Lemma 2.17, (47)
and Lemma 2.3 (with σ = σN ln 2/2 and η = 3/4), there is δ ∈ (0, ζ) such that
σN √
L∆ Ve (yε ) > ln 2
2
for all yε = (ε, y 0 ) ∈ Bδ ∩ B2+ .
Proof of Theorem 2.1. The claim follows from Theorems 2.4 (for the case N = 1)
and Theorem 2.19 (for the case N ≥ 2).
3. A new barrier via the Kelvin transform. Let R > 0 and x0 ∈ RN . The
inversion of a point x ∈ RN with respect to the sphere SR (x0 ) is given by
x − x0
x∗ := x0 + R2 , x 6= x0 , (48)
|x − x0 |2
and the Kelvin transform of u is
u∗ (x∗ ) := |x∗ − x0 |−N u(x), x∗ 6= x0 . (49)
22 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
|z|−2N
Z
= cN ∗ N
dz + R1 (x) (53)
Bρ (P )\Ω∗ |z − x|
|z|−N |x∗ |N
Z
= cN ∗ N
dz + R1 (x).
Bρ (P )\Ω∗ |z − x |
On the other hand, let A2 = [B1 (x∗ )\(Bσ (x∗ )∪Ω)]∪[Ω\B1 (x∗ )] (with σ as in (52))
and Z
∗ 1
R2 (x ) = cN dz.
A2 |z − x∗ |N
OPTIMAL BOUNDARY REGULARITY FOR THE LOGARITHMIC LAPLACIAN 23
By (59), it is clear that u∗ ∈ L∞ (Ω∗ ). Moreover, by Lemma 3.1, (59), and (10), it
also follows that L∆ u∗ ∈ L∞ (Ω∗ ), as claimed.
To prove iii), let u ∈ H(Ω) and recall that we have assumed that x0 = 0,
R = 1, and that (59) holds. Note that, in this setting, Ω∗ ⊂ B1 , B2 (y) ⊂ B3
for all y ∈ Ω∗ , and B2 (y) ⊂ RN \Ω∗ for all y ∈ RN \B3 . To ease notation, let
∗ ∗
(y))2
S = S(x, y) := (u (x)−u
|x−y|N
, then, using that u∗ = 0 in RN \Ω∗ ,
Z Z Z Z Z Z
S dxdy = S dxdy + S dxdy
RN B2 (y) Ω∗ B (y) RN \Ω∗ B2 (y)
Z Z 2 Z Z Z Z
= S dxdy + S dxdy + S dxdy.
Ω∗ Ω∗ Ω∗ B2 (y)\Ω∗ B3 \Ω∗ B2 (y)
Note that, by (59), there is δ > 0 such that Bδ ⊂ (Ω∗ )c and therefore there is
some M > 0 such that
(B3 \Ω∗ )∗ ⊂ A\Ω, A := BM \B 31 .
Using that Ω is bounded, (59), and that (|ξ|N u(ξ) − |z|N u(z))2 ≤ 2(|ξ|2N (u(ξ) −
u(z))2 + 2(|ξ|N − |z|N )2 u(z)2 , we have that
2
|ξ|N u(ξ) − |z|N u(z)
Z Z
|ξ|−N |z|−N dξdz
Ω Ω |ξ − z|N
(u(ξ) − u(z))2 N −N (|ξ|N − |z|N )2
Z Z
≤2 |ξ| |z| + u(z)2 |ξ|−N |z|−N dξdz
Ω Ω |ξ − z|N |ξ − z|N
(u(ξ) − u(z))2
Z Z Z
≤C dξdz + C 1 u(z)2 dz (61)
Ω Ω |ξ − z|N Ω
OPTIMAL BOUNDARY REGULARITY FOR THE LOGARITHMIC LAPLACIAN 25
Moreover, using (59) and that Ω is bounded, we find some C2 = C2 (Ω) > 0 such
that
u(z)2 |z|N
Z Z Z Z
N |ξ|N
dξdz < C 2 u(z)2
|ξ − z|−N dξdz
Ω A\Ω |ξ − z| Ω A\Ω
Z Z
C2 2
= u(z) κΩ (z)dz + C2 |A| u(z)2 dz, (62)
cN Ω Ω
where
R κΩ is the so-called killing measure (see [10, eq. (3.5)]) given by κΩ (z) :=
cN B1 (z)\Ω |z − y|−N dy for z ∈ Ω. By the logarithmic boundary Hardy inequality
(see [10, Remark 4.3 and Corollary A.2]), we have that Ω u(z)2 κΩ (z)dz < ∞. Claim
R
Let κ : RN → RN be the Kelvin transform κ(x) := |x|x 2 , that is, the point inversion
with respect to the sphere S1 . For x 6= 0, let v(x) := w(κ(x))
e and let D be as in
the statement. It is not hard to see that κ(Ω)
e = D and that
e ∩ {x ∈ RN : x1 = 1}) = ∂D ∩ ∂B 1 ( 1 e1 ),
κ(∂ Ω 2 2
26 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
namely, that the flat part of the boundary of Ω e goes to the curved part of the
boundary of D through κ (see Figure 2). Using Proposition 1.3, it follows that
v ∈ C ∞ (D) ∩ H(D) is a solution of L∆ v = f in D for some f ∈ L∞ (D). A
1
simple calculation using Lemma 1.5 and (50) yields that v ∈ L 2 (RN ) and that (63)
holds.
We close this subsection with the following result on weak solutions under the
Kelvin transform.
Lemma 3.5. Let Ω ⊂ RN be as in Proposition 1.3 ii). Let u be a weak solution of
L∆ u = f in Ω, u=0 in RN \ Ω,
Then u∗ is a weak solution of
L∆ u∗ = f¯ in Ω∗ , u∗ = 0 in RN \ Ω∗ ,
where
|z|−N − |y ∗ |−N
Z
f¯(y) := f (y ∗ )|y|−N + cN u(y ∗ )|y|−2N dz
Ω |z − y ∗ |N
+ u(y ∗ )(hΩ∗ (y) − hΩ (y ∗ ))|y|−N , y ∈ Ω∗ .
for y ∈ Ω∗ . Thus, using that u∗ (y) = |y|−N u(y ∗ ) and ψ(y) = |y|−N ϕ(y ∗ ),
|z|−N − |y ∗ |−N
Z Z
E(u∗ , ψ) = u∗ (y)|y|−N L∆ ϕ(y ∗ ) + cN u∗ (y)|y|−2N ϕ(y ∗ ) dz
Ω∗ Ω |z − y ∗ |N
+ u∗ (y)(hΩ∗ (y) − hΩ (y ∗ ))|y|−N ϕ(y ∗ ) dy
|z|−N − |y ∗ |−N
Z Z
= u(y ∗ )|y|−2N L∆ ϕ(y ∗ ) + cN u(y ∗ )|y|−2N ψ(y) dz
Ω∗ Ω |z − y ∗ |N
+ u(y ∗ )(hΩ∗ (y) − hΩ (y ∗ ))|y|−N ψ(y ∗ ) dy.
Using that u is a weak solution and the change of variables y = x∗ , we obtain that
Z Z
u(y ∗ )|y|−2N L∆ ϕ(y ∗ ) dy = u(x)L∆ ϕ(x) dx
Ω∗ Ω
f (y ∗ )
Z Z
= E(u, ϕ) = f (x)ϕ(x) dx = N
ψ(y) dy.
Ω Ω∗ |y|
5. Applications.
5.1. Optimal regularity for the torsion function in a small ball. In this
section, we show Theorem 1.2. We begin with some existence and qualitative prop-
erties for the torsion function.
N N
Proposition 5.1. Let N ≥ 1, r > 0 be such that |Br | < 2N e 2 (Ψ( 2 )−γ) |B1 |, and
consider the following problem
L∆ τ = 1 in Br , τ =0 in RN \Br . (66)
There is a unique classical positive solution τ to (66). In particular, τ is radially
symmetric and continuous in RN .
Proof. Under the assumption on r, by Theorem 4.1, we have that L∆ satisfies the
maximum principle in Br . By [6, Theorem 1.1], there is a classical solution of (66),
namely, (66) holds pointwisely. Since L∆ satisfies the maximum principle in Br ,
we have that τ ≥ 0 in Br . Moreover, τ > 0 in Br , because, if τ (y0 ) = 0 for some
y0 ∈ Ω, then we would have that
Z
τ (y)
1 = L∆ τ (y0 ) = −cN N
dy < 0,
B |y 0 − y|
which is absurd. Finally, τ is radially symmetric by the uniqueness of the
solution.
5.1.1. 1D case. We show the one-dimensional case first.
1 1
Theorem 5.2. Let r > 0 be such that r < 2e 2 (Ψ( 2 )−γ) ≈ 0.561459 and let τ be the
torsion function of the interval (0, r), namely, the unique weak solution of
L∆ τ = 1 in (0, r), τ =0 in R \ (0, r).
Then there is c ∈ (0, 1) such that
1
` 2 (δ(x)) 1
≥ τ (x) ≥ c ` 2 (δ(x)) for x ∈ (0, r),
c
where δ(x) = dist(x, ∂(0, r)) = dist(x, {0, r}) = min{x, r − x}.
Proof. The upper bound follows from Theorem 1.1. For the lower bound, let u be
given by (26). Then L∆ u ∈ L∞ (Ω) with Ω = (0, 2), by Theorem 2.4. Let r > 0 be
as in the statement, ε := 41 r, Ω
e := (0, 2ε) ⊂ (0, r), and let u e → R be given by
e:Ω
x ∞ e
u
e(x) := u( ε ). By Lemma A.3, L∆ u e ∈ L (Ω).
By Proposition 5.1, τ > 0 in [0, r) and τ ∈ C(R). Let U := τ − kL∆1uek∞ u e, then
L∆ U ≥ 0 in Ω and U ≥ 0 on R\Ω. By Theorem 4.1, we obtain that U ≥ 0 in R,
e e
namely
C 1
x 1
τ (x) ≥ C u
e(x) = p x
= C` 2 ≥ C1 ` 2 (x) for x ∈ (0, 10 )
− ln 2 2
for some constants C, C1 > 0, by Lemma 1.5. One can argue similarly in a neigh-
1
borhood of r to obtain that τ (x) ≥ C2 ` 2 (r − x) for x ∈ (r − 10 , r) and for some
constant C2 > 0. Let I(ε) := ( 10 , r − 10 ). Since τ > 0 in (0, r), then
1 inf I(ε) τ
τ ≥ C3 ` 2 (δ(x)) for x ∈ I(ε), C3 := 1 .
supI(ε) ` 2 (δ(x))
The lower bound now follows with c := min{C1 , C2 , C3 }. This ends the proof.
OPTIMAL BOUNDARY REGULARITY FOR THE LOGARITHMIC LAPLACIAN 29
Since τ is radially symmetric, then the lower bound in (8) follows. The lower
estimate in (9) now follows from (68) (with a proof by contradiction, for instance).
On the other hand, the upper bound in (8) and in (9) holds by Theorem 1.1.
5.2. A Hopf-type lemma for the logarithmic Laplacian. For a bounded open
set Ω, the space V(Ω) is defined in (11). We say that a function v ∈ V(Ω) solves
weakly that L∆ v ≥ 0 in Ω if
EL (v, ϕ) ≥ 0 for all ϕ ∈ Cc∞ (Ω) with ϕ ≥ 0 in Ω.
We remark that EL (v, ϕ) < ∞ for v ∈ V(Ω) and ϕ ∈ Cc∞ (Ω) by [10, Lemma 4.4].
Recall that η denotes the outer unit normal vector field along ∂Ω. We are ready to
show Theorem 1.4.
Theorem 5.4. For every µ > 0 there is a unique (up to a sign) least-energy weak
solution u ∈ H(Ω) of (72) which is a global minimizer of the energy functional
Z
1 µ
v 2 ln(v 2 ) − 1 dx. (73)
J0 : H(Ω) → R, J0 (v) := EL (v, v) + I(v), I(v) :=
2 4 Ω
1 1
Moreover, 0 < |u(x)| ≤ (R2 e 2 −ρN ) µ for x ∈ Ω, where R := 2 diam(Ω) and ρN is an
explicit constant given in (14). Furthermore, u ∈ C(RN ), and there are α ∈ (0, 1)
and C > 0 such that
|u(x) − u(y)|
sup < C. (74)
x,y∈R N `α (|x − y|)
x6=y
In this theorem, the uniqueness (up to a sign) of the least energy solution is
shown using a convexity-by-paths argument as in [4, Section 6]. In particular, the
following is shown: Given u and v in H(Ω) such that u2 6= v 2 , let
1
θ(t, u, v) := ((1 − t)u2 + tv 2 ) 2 , (75)
then
the function t 7→ J0 (θ(t, u, v)) is strictly convex in [0, 1], (76)
Since a strictly convex function cannot have two global minimizers, (76) immediately
yields the uniqueness of least-energy solutions.
Now we use (76) and Theorem 1.4 to yield the uniqueness of positive solutions
(which has to be the positive least-energy solution obtained in Theorem 5.4).
Theorem 5.5. For every µ > 0 there is only one positive and bounded solution
of (72).
Proof. Let A be the set of positive and bounded critical points of J0 . By The-
orem 5.4 we know that A is nonempty. Let u, v ∈ A. Since u, v ∈ L∞ (Ω), it
follows that ln |u|u, ln |v|v ∈ L∞ (Ω), and, by [10, Theorem 1.11], we have that
u, v ∈ C(Ω). Then there is ε > 0 such that −µ ln |u|u > 0 and −µ ln |v|v > 0 in
Ωε := {x ∈ Ω : dist(x, ∂Ω) < ε}. Since u and v are continuous weak solutions
of (72), we have in particular that u and v solve weakly
L∆ u ≥ 0, L∆ v ≥ 0 in Ωε , u ≥ 0, v ≥ 0 in RN \Ωε .
By Theorem 1.4, there is c ∈ (0, 1) such that
u(x0 − tη(x0 )) u(x0 − tη(x0 ))
c−1 > lim sup 1 ≥ lim inf 1 > c,
t→0+ ` 2 (t) t→0 +
` 2 (t)
v(x0 − tη(x0 )) v(x0 − tη(x0 ))
c−1 > lim sup 1 ≥ lim inf 1 > c,
` 2 (t)
t→0+ t→0+ ` 2 (t)
for all x0 ∈ ∂Ω. Then, by Corollary 5.3, we obtain that u and v are comparable,
namely, that there is M > 1 such that
v(x)
M> > M −1 for x ∈ Ω. (77)
u(x)
v
Let θ be as in (75). We now argue as in [4, Theorem 6.1]. Let w := u χΩ ,
2
ξ −1
z(ξ, t) := 1 , and let χΩ denote the characteristic function of Ω, then
(1−t+tξ 2 ) 2 +1
v 2 − u2 w2 − 1
= =u 1 = uz(w, t).
θ(t) + u (1 − t + tw2 ) 2 + 1
By (77), for x, y ∈ Ω,
u(x)z(w(x), t) − u(y)z(w(y), t)
= (u(x) − u(y))z(w(x), t) − u(y)(z(w(y), t) − z(w(x), t))
and, by the Mean-Value Theorem,
v(x)
u(y)|z(w(x), t) − z(w(y), t)| ≤ C1 u(y)|w(x) − w(y)| = C1 u(y) − v(y)
u(x)
v(x)
= C1 v(x) − v(y) + (u(y) − u(x))
u(x)
≤ C2 (|v(x) − v(y)| + |u(y) − u(x)|),
where C1 := sup(k,t)∈[0,M ]×[0,1] |∂ξ z(k, t)| < ∞ and C2 := C1 + M .
On the other hand, if x ∈ Ω and y ∈ RN \Ω, then u(y) = 0 and
|u(x)z(w(x), t) − u(y)z(w(y), t)| ≤ C1 M |u(x)| = C1 M |u(x) − u(y)|.
Then there is C3 > 0 such that
2
θ(t) − θ(0)
t
Z Z | θ(t)−θ(0)
t (x) − θ(t)−θ(0)
t (y)|2
= dy dx
RN B1 (x) |x − y|N
!
|u(x) − u(y)|2 |v(x) − v(y)|2
Z Z Z Z
≤ C3 dy dx + dy dx
RN B1 (x) |x − y|N RN B1 (x) |x − y|N
= C3 (kuk2 + kvk2 ).
This, together with (76), guarantees that all the assumptions of [4, Theorem 1.1]
are satisfied. Then, this result implies that A has at most one element, and this
ends the proof.
where we used [10, Theorem 1.1] to justify the first and last equalities.
The next result is an integration by parts formula under slightly weaker assump-
tions than those in [10, equation (3.11)].
Lemma A.2 (Integration by parts). Let Ω ⊂ RN be a bounded Lipschitz set and
let w ∈ H(Ω) be such that L∆ w ∈ L∞ (Ω). Then,
Z
[L∆ w] v dx = EL (w, v) for all v ∈ H(Ω).
RN
Proof. A similar result for uniformly Dini continuous functions can be found in [10,
equation (3.11)], and the same arguments can be extended for w as in the statement.
We give a proof for completeness. Let k : RN \{0} → R and j : RN → R be given
by
k(z) := cN 1B1 (z)|z|−N and j(z) := cN 1RN \B1 (z)|z|−N .
Then, for x ∈ Ω,
Z
L∆ w(x) = (w(x) − w(y))k(x − y)dy − [j ∗ w](x) + ρN w(x)
RN
and, by a standard argument using a change of variables,
Z
[L∆ w] v dx
RN
Z Z
= (w(x) − w(y))k(x − y)dy − [j ∗ w](x) + ρN w(x) v(x) dx
RN N
Z ZR Z
1
= (w(x) − w(y))(v(x) − v(y))k(x − y) dx dy − (j ∗ w − ρN w) v dx
2 RN RN RN
= EL (w, v)
for all v ∈ H(Ω), as claimed.
Lemma A.3 (On weak solutions). Let u ∈ H(Ω) ∩ L∞ (Ω) be a weak solution of
L∆ u = f in Ω, u = 0 on RN \Ω for some f ∈ L∞ (Ω). Then, for any λ > 0, uλ is a
weak solution of L∆ uλ = feλ in λΩ, uλ = 0 on RN \λΩ, with feλ := fλ + ln(λ−2 )uλ ∈
L∞ (Ω).
Proof. Let ϕ ∈ Cc∞ (Ω). Using Lemmas A.1 and A.2,
Z
EL (uλ , ϕλ ) = uλ L∆ ϕλ dx
N
ZR
= u(λ−1 x)L∆ ϕ(λ−1 x) + ln(λ−2 )u(λ−1 x)ϕ(λ−1 x) dx
R N
Z
= λN uL∆ ϕ + ln(λ−2 )uϕ dx
RN
Z
N −2
=λ EL (u, ϕ) + ln(λ ) uϕ dx
RN
Z
= (fλ + ln(λ−2 )uλ )ϕλ dx.
RN
Now, a scaling property for pointwise solutions easily follows from the previous
result.
34 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
Lemma A.4 (On pointwise solutions). Let w and Ω be as in Lemma A.2. Moreover,
assume that w ∈ L∞ (Ω) and let f (x) := L∆ w(x) for x ∈ Ω. Then, for any
λ > 0, wλ is a solution of L∆ wλ = feλ in Ωλ , wλ = 0 on RN \Ωλ , with feλ :=
fλ + ln(λ−2 )wλ ∈ L∞ (Ωλ ).
Proof. Let λ > 0, w, and Ω as in the statement, ϕ ∈ Cc∞ (Ωλ ), and let ψ(x) :=
ϕ(λx). Then, by Lemmas A.1, A.2, and a change of variables,
Z Z Z Z
λ−N fλ ϕ dx = f ψ dx = L∆ wψ dx = EL (w, ψ) = wL∆ ψ dx
Ωλ Ω Ω Ω
Z
= w(x)(L∆ ϕ(λx) + ln(λ2 )ϕ(λx)) dx
Ω
Z
−N
=λ wλ (x)(L∆ ϕ(x) + ln(λ2 )ϕ(x)) dx
Ωλ
Z
−N
=λ (L∆ wλ (x) + ln(λ2 )wλ (x))ϕ(x) dx.
Ωλ
R1 ωu,x,E (r)
and u is said to be Dini continuous at x if 0 r dr < ∞. Let
|u(x)|
Z
L10 (RN ) := u ∈ L1loc (RN ) : dx < ∞ .
RN (1 + |x|)N
Lemma A.5. Let u, v ∈ L10 (RN ) be such that u and v are Dini continuous functions
at some x ∈ RN and v ∈ L∞ (RN ). Then L∆ [uv](x) is well defined (in the sense of
formula (2)) and
L∆ [uv](x) = u(x)L∆ v(x) + v(x)L∆ u(x) − I(u, v)(x), (78)
where
I(u, v)(x)
(u(x) − u(y))(v(x) − v(y))
Z
= cN dy
B1 (x) |x − y|N
u(y)v(y) − u(x)v(y) − u(y)v(x)
Z
+ cN dy + ρN u(x)v(x). (79)
RN \B1 (x) |x − y|N
Proof. Let x ∈ Ω be as in the statement. First we show that the right hand side
of (78) is well defined. The first two terms, u(x)L∆ v(x) and v(x)L∆ u(x) are well
defined by our assumptions on u and v, see [10, Proposition 2.2]. On the other
hand, since v ∈ L∞ (RN ) and u is Dini continuous at x,
(u(x) − u(y))(v(x) − v(y)) (u(x) − u(z + x))(v(x) − v(z + x))
Z Z
N
dy = dz
B1 (x) |x − y| B1 |z|N
OPTIMAL BOUNDARY REGULARITY FOR THE LOGARITHMIC LAPLACIAN 35
1
|u(x) − u(x + z)|
Z Z
ωu,x,B1 (r)
≤ 2kvk L∞ dz ≤ 2kvk∞ |SN −1 | dr < ∞.
B1 |z|N 0 r
∞ N
Furthermore, since v ∈ L (R ) and u ∈ L10 (R),
there is a constant Cx > 0 only
depending on x such that
|u(y)|
Z Z
u(y)v(y)
N
dy ≤ kvk L ∞ dy
N
R \B1 (x) |x − y| N
R \B1 (x) |x − y|N
|u(y)|
Z
≤ Cx kvkL ∞ dy < ∞.
RN (1 + |y|)N
Similarly, RN \B1 (x) u(x)v(y) dy and RN \B1 (x) u(y)v(x)
R R
|x−y|N |x−y|N
dy are also finite. Then, since
L∆ [uv](x)
u(x)v(x) − u(y)v(y)
Z Z
u(y)v(y)
= cN dy − cN dy + ρN u(x)v(x),
B1 (x) |x − y|N RN \B1 (x) |x − y|N
identity (78) follows by noting that
u(x)v(x) − u(y)v(y)
Z
cN dy
B1 (x) |x − y|N
v(x) − v(y) (u(x) − u(y))v(y)
Z Z
= u(x) cN N
dy + cN dy
B1 (x) |x − y| B1 (x) |x − y|N
v(x) − v(y) u(x) − u(y)
Z Z
= u(x) cN N
dy + v(x) cN N
dy
B1 (x) |x − y| B1 (x) |x − y|
(u(x) − u(y))(v(x) − v(y))
Z
− cN dy.
B1 (x) |x − y|N
REFERENCES
[1] N. Abatangelo, S. Dipierro, M. M. Fall, S. Jarohs and A. Saldaña, Positive powers of the
Laplacian in the half-space under Dirichlet boundary conditions, Discrete and Continuous
Dynamical Systems, 39 (2019), 1205-1235.
[2] F. Angeles and A. Saldaña, Small order limit of fractional Dirichlet sublinear-type problems,
Fract. Calc. Appl. Anal., 26 (2023), 1594-1631.
[3] K. Bogdan and T. Ẑak, On Kelvin transformation, J. Theor Probab, 19 (2006), 89-120.
[4] D. Bonheure, J. Földes, E. Moreira dos Santos, A. Saldaña and H. Tavares, Paths to unique-
ness of critical points and applications to partial differential equations, Trans. Amer. Math.
Soc., 370 (2018), 7081-7127.
36 V. HERNÁNDEZ-SANTAMARÍA, L. F. LÓPEZ RÍOS AND A. SALDAÑA
[5] C. Bucur, Some observations on the Green function for the ball in the fractional Laplace
framework, Commun. Pure Appl. Anal., 15 (2016), 657-699.
[6] H. A. Chang-Lara and A. Saldaña, Classical solutions to integral equations with zero order
kernels, Mathematische Annalen, 389 (2024), 1463–1515.
[7] H. Chen, Taylor’s expansions of Riesz convolution and the fractional Laplacians with respect
to the order, arXiv preprint, arXiv:2307.06198, 2023.
[8] H. Chen, D. Hauer and T. Weth, An extension problem for the logarithmic Laplacian, arXiv
preprint, arXiv:2312.15689, 2023.
[9] H. Chen and L. Véron, The cauchy problem associated to the logarithmic laplacian with an
application to the fundamental solution, Journal of Functional Analysis, 287 (2024), 110470.
[10] H. Chen and T. Weth, The Dirichlet problem for the logarithmic Laplacian, Comm. Partial
Differential Equations, 44 (2019), 1100-1139.
[11] E. Correa and A. De Pablo, Nonlocal operators of order near zero, Journal of Mathematical
Analysis and Applications, 461 (2018), 837-867.
[12] J. Dávila, L. López Rı́os and Y. Sire, Bubbling solutions for nonlocal elliptic problems, Revista
Matemática Iberoamericana, 33 (2017), 509-546.
[13] A. DelaTorre and E. Parini, Uniqueness of least energy solutions to the fractional Lane-Emden
equation in the ball, arXiv preprint, arXiv:2310.02228, 2024.
[14] A. Dieb, I. Ianni and A. Saldaña, Uniqueness and nondegeneracy for Dirichlet fractional
problems in bounded domains via asymptotic methods, Nonlinear Analysis, 236 (2023),
113354.
[15] A. Dieb, I. Ianni and A. Saldaña, Uniqueness and nondegeneracy of least-energy solutions to
fractional Dirichlet problems, arXiv preprint, arXiv:2310.01214, 2023.
[16] M. M. Fall and S. Jarohs, Overdetermined problems with fractional Laplacian, ESAIM: Con-
trol, Optimisation and Calculus of Variations, 21 (2015), 924-938.
[17] M. M. Fall and T. Weth, Nonradial nondegeneracy and uniqueness of positive solutions to a
class of fractional semilinear equations, arXiv preprint, arXiv:2310.10577, 2023.
[18] M. M. Fall and T. Weth, Second radial eigenfunctions to a fractional Dirichlet problem and
uniqueness for a semilinear equation, arXiv preprint, arXiv:2405.02120, 2024.
[19] P. A. Feulefack, The fractional logarithmic Schrodinger operator: Properties and functional
spaces, arXiv preprint, arXiv:2310.02481, 2023.
[20] P. A. Feulefack, The logarithmic Schrödinger operator and associated Dirichlet problems,
Journal of Mathematical Analysis and Applications, 517 (2023), 126656.
[21] P. A. Feulefack and S. Jarohs, Nonlocal operators of small order, Ann. Mat. Pura Appl. (4),
202 (2023), 1501-1529.
[22] P. A. Feulefack, S. Jarohs and T. Weth, Small order asymptotics of the Dirichlet eigenvalue
problem for the fractional Laplacian, Journal of Fourier Analysis and Applications, 28 (2022),
Paper No. 18, 44 pp.
[23] V. Hernández-Santamarı́a, S. Jarohs, A. Saldaña and L. Sinsch, FEM for 1D-problems involv-
ing the logarithmic Laplacian: Error estimates and numerical implementation, arXiv preprint,
arXiv:2311.13079, 2023.
[24] V. Hernández Santamarı́a and A. Saldaña, Small order asymptotics for nonlinear fractional
problems, Calculus of Variations and Partial Differential Equations, 61 (2022), 1-26.
[25] S. Jarohs, A. Saldaña and T. Weth, A new look at the fractional Poisson problem via the
logarithmic Laplacian, J. Funct. Anal., 279 (2020), 108732, 50 pp.
[26] S. Jarohs, A. Saldaña and T. Weth, Differentiability of the nonlocal-to-local transition in
fractional Poisson problems, arXiv preprint, arXiv:2311.18476, 2023.
[27] M. Kassmann and A. Mimica, Intrinsic scaling properties for nonlocal operators, J. Eur.
Math. Soc. (JEMS), 19 (2017), 983-1011.
[28] P. Kim and A. Mimica, Green function estimates for subordinate Brownian motions: Stable
and beyond, Transactions of the American Mathematical Society, 366 (2014), 4383-4422.
[29] A. Laptev and T. Weth, Spectral properties of the logarithmic Laplacian, Anal. Math. Phys.,
11 (2021), Paper No. 133, 24 pp.
[30] X. Ros-Oton and J. Serra, The Dirichlet problem for the fractional Laplacian: Regularity up
to the boundary, Journal de Mathématiques Pures et Appliquées, 101 (2014), 275-302.
Received February 2024; revised June 2024; early access June 2024.