
Citation: Isabeau Birindelli, Giulio Galise. Allen-Cahn equation for the truncated Laplacian: Unusual phenomena[J]. Mathematics in Engineering, 2020, 2(4): 722-733. doi: 10.3934/mine.2020034
[1] | Juan-Carlos Felipe-Navarro, Tomás Sanz-Perela . Semilinear integro-differential equations, Ⅱ: one-dimensional and saddle-shaped solutions to the Allen-Cahn equation. Mathematics in Engineering, 2021, 3(5): 1-36. doi: 10.3934/mine.2021037 |
[2] | Francesca G. Alessio, Piero Montecchiari . Gradient Lagrangian systems and semilinear PDE. Mathematics in Engineering, 2021, 3(6): 1-28. doi: 10.3934/mine.2021044 |
[3] | David Cruz-Uribe, Michael Penrod, Scott Rodney . Poincaré inequalities and Neumann problems for the variable exponent setting. Mathematics in Engineering, 2022, 4(5): 1-22. doi: 10.3934/mine.2022036 |
[4] | Antonio Vitolo . Singular elliptic equations with directional diffusion. Mathematics in Engineering, 2021, 3(3): 1-16. doi: 10.3934/mine.2021027 |
[5] | Luigi Montoro, Berardino Sciunzi . Qualitative properties of solutions to the Dirichlet problem for a Laplace equation involving the Hardy potential with possibly boundary singularity. Mathematics in Engineering, 2023, 5(1): 1-16. doi: 10.3934/mine.2023017 |
[6] | Italo Capuzzo Dolcetta . The weak maximum principle for degenerate elliptic equations: unbounded domains and systems. Mathematics in Engineering, 2020, 2(4): 772-786. doi: 10.3934/mine.2020036 |
[7] | Edgard A. Pimentel, Miguel Walker . Potential estimates for fully nonlinear elliptic equations with bounded ingredients. Mathematics in Engineering, 2023, 5(3): 1-16. doi: 10.3934/mine.2023063 |
[8] | Marco Cirant, Kevin R. Payne . Comparison principles for viscosity solutions of elliptic branches of fully nonlinear equations independent of the gradient. Mathematics in Engineering, 2021, 3(4): 1-45. doi: 10.3934/mine.2021030 |
[9] | Filippo Gazzola, Gianmarco Sperone . Remarks on radial symmetry and monotonicity for solutions of semilinear higher order elliptic equations. Mathematics in Engineering, 2022, 4(5): 1-24. doi: 10.3934/mine.2022040 |
[10] | Yuzhe Zhu . Propagation of smallness for solutions of elliptic equations in the plane. Mathematics in Engineering, 2025, 7(1): 1-12. doi: 10.3934/mine.2025001 |
To Sandro Salsa, a fine mathematician with a taste for life.
In this paper we consider some entire solutions of a degenerate elliptic equation with non linear forcing term, where the behaviour of the solutions is quite different from the uniformly elliptic case.
We begin by recalling the linear uniformly elliptic model problem we have in mind in order to underline the differences with the degenerate elliptic case that we shall treat here. Consider the entire solutions of the Allen-Cahn equation
Δu+u−u3=0in RN | (1.1) |
where N≥2.
Writing the variable as x=(x′,xN), with xN∈R, we are interested in solutions that satisfy the following conditions:
|u|<1 and limxN→±∞u(x′,xN)=±1 uniformly in x′∈RN−1. | (1.2) |
Any solution of (1.1) satisfying (1.2) is necessarily a function of xN, i.e. it is independent of x′. This was referred to as Gibbons conjecture and it has been proved by several authors [3, 5, 18, 19] and then it has been generalized to different contexts [15,20,23]. It is well known that this conjecture is somehow related to a famous conjecture of De Giorgi on the 1 dimensionality of monotone solutions in low dimension, see for its proof [4,21] in the 2 dimensional case, [2] for the 3 dimensional case and for higher cases with some mild further assumptions, the work of Savin [24]. For a counterexample in dimensions N≥9, providing a sharpness of the conjecture, was constructed by Del Pino, Kowalczyk and Wei in [17]
The function f(u)=u−u3 can be replaced by a much more general class of functions and the result is still valid. On the other hand, if instead of the Laplacian one considers some degenerate elliptic operators, there are few results, let us mention e.g., the works in the Heisenberg group and in Carnot groups [14,16].
In this work we shall mainly focus on equations whose leading term is the operator P−k which is sometimes referred to as the truncated Laplacian. The operator P−k is defined, for any N×N symmetric matrix X, by the partial sum
P−k(X)=λ1(X)+…+λk(X) |
of the ordered eigenvalues λ1≤…≤λN of X. We shall consider solutions of the equation
P−k(D2u)+f(u)=0in RN, | (1.3) |
for a general class of functions f modelled on f(u)=u−u3. Clearly P−k(D2u) corresponds to the Laplacian when k=N, hence in the whole paper we shall suppose that k=1,…,N−1.
These operators are degenerate elliptic in the sense that
X≤Y⇒P−k(X)≤P−k(Y). |
In order to emphasize the strong degeneracy of these operators, let us mention that for any matrix X there exists M≥0, not identically zero, such that
P−k(X)=P−k(X+M). |
It is immediate to see that, for example, one can take M=vN⊗vN where vN is an eigenvector corresponding to the largest eigenvalue of X.
In previous works we have, together with Hitoshi Ishii and/or Fabiana Leoni [9,10,11,13], encountered a certain number of surprising results, related to these degenerate operators. It would be too long to recall them all here but we shall briefly recall some of those closer to the results in this paper.
The classical Liouville result states that any harmonic function which is bounded from below is a constant. This is not true for bounded from below entire solutions of P−k(D2u)=0.
Similarly, concerning the semi-linear Liouville theorem, the existence of entire non negative solutions of F(D2u)+up=0 is quite different if F(D2u)=Δu or if F(D2u)=P−k(D2u). Indeed, for the Laplacian there is a threshold for p between existence and non existence. This threshold is different for solutions and supersolutions and it depends on the dimension of the space. Instead, for solutions of the equation
P−k(D2u)+up=0 in RN | (1.4) |
the following hold:
1). For any p>0 there exist nonnegative viscosity solutions u≢0;
2). For any p≥1 there exist positive classical solutions;
3). For p<1 there are no positive viscosity supersolutions of (1.4).
Interestingly, if one considers instead non positive solutions, the non existence results are very similar to those of the Laplacian in Rk. These Liouville theorems were proved in [13].
Hence this lead us to wonder what happens to bounded entire solutions of the equation
P−k(D2u)+u−u3=0 | (1.5) |
which satisfy |u|<1 and which, a priori, may change sign. Can one expect solutions to be one dimensional? Precisely, can we extend Gibbons conjecture to this degenerate case?
In order to answer these questions the first step is to study the one dimensional solutions, which is what we do in Section 2. Interestingly the results are completely different from the uniformly elliptic case and this leads to different conjectures. Precisely, we consider only one dimensional solutions u(x′,xN)=v(xN) of (1.5) that satisfy |u|<1. The results can be summarized in the following way:
1). The only classical solution is u≡0;
2). Any viscosity subsolution is non negative and there exists a non trivial viscosity solution that satisfies
limxN→−∞v(xN)=0 and limxN→∞v(xN)=1. | (1.6) |
3). There are no solutions that are strictly monotone or positive
Observe that, even though u3−u is the derivative of the double well potential F(u)=14(1−u2)2, the lack of ellipticity does not allow the solutions to go from −1 to 1. This is the first surprising result.
Hence Gibbons conjecture should be reformulated in the following way:
Question 1. Is it true that if u is a solution of (1.5), |u|<1, satisfying
limxN→−∞u(x′,xN)=−1 and limxN→∞u(x′,xN)=1 uniformly in x′∈RN−1 |
then u is 1 dimensional?
If the answer was positive, this would imply that there are no solutions of (1.5) that satisfies (1.2), since such one dimensional solutions don't exist in view of point 2 above. Or, equivalently, if such a solution of (1.5) exists then the answer to Question 1 is negative.
Question 2. Is it true that if u is a solution of (1.5), |u|<1, satisfying
limxN→−∞u(x′,xN)=0 and limxN→∞u(x′,xN)=1 uniformly in x′∈RN−1 |
then u is 1 dimensional?
This is more similar in nature to the uniformly elliptic case. Nonetheless, classical proofs of these symmetry results rely heavily on the strong maximum principle, or strong comparison principle, and on the sliding method and the moving plane method [7], which in general don't hold for the truncated Laplacian (see [9,12]). And in particular they are not true here since we construct ordered solutions that touch but don't coincide.
Another remark we wish to make is that, even though the solutions we consider are one dimensional, since they are viscosity solutions, the test functions are not necessarily one dimensional. Hence the proofs are not of ODE type.
Other surprising results concern Liouville type theorems for (1.5) i.e., existence of bounded entire solutions bounded without requiring conditions at infinity. Aronson and Weinberger in [1] and more explicitly, Berestycki, Hamel and Nadirashvili in [6] have proved the following Liouville type result:
If v is a bounded non negative classical solution of
Δu+u−u3=0 in RN |
then either u≡0 or u≡1.
Once again this result fails if one replaces the Laplacian with the truncated Laplacian.
Indeed we prove that there exists infinitely many bounded non negative smooth solutions of
P−k(D2u)+f(u)=0, in RN |
for a general class of nonlinearities that includes f(u)=u−u3. This is done by constructing infinitely many radial solutions of P−k(D2u)+f(u)=0 which are positive in RN but tend to zero at infinity.
Finally in Section 4 we show a different surprising phenomena related to the so called principal eigenvalue of P−k. This is somehow different in nature but we believe that it sheds some light to these extremal degenerate operators.
We consider one dimensional viscosity solutions u, i.e., u(x)=v(xN) for x=(x1,…,xN), of the problem
{P−k(D2u)+u−u3=0in RN|u|<1. | (2.1) |
The main result is the following.
Proposition 1. Concerning problem (2.1), the following hold:
i) If u∈USC(RN) is a viscosity one dimensional subsolution then u≥0.
ii) The only classical one dimensional solution is u≡0.
iii) There exist nontrivial viscosity one dimensional solutions, e.g., u(x)=v(xN), satisfying either
limxN→−∞v(xN)=0 andlimxN→∞v(xN)=1. | (2.2) |
or
limxN→±∞v(xN)=1. | (2.3) |
iv) There are no positive viscosity one dimensional supersolutions.
v) If u≥0 is a viscosity one dimensional supersolution e.g., u(x)=v(xN) and it is nondecreasing in the xN-direction then there exists t0∈R such that
u=0 in X0={x∈RN:xN≤t0}. |
Remark 1. A consequence of i) and v) is that there are no viscosity one dimensional solutions increasing in the xN-direction.
Proof. i) Fix ˆx=(ˆx1,…,ˆxN)∈RN and, for α>0, let
max¯B1(ˆx)[u(x)−α(xN−ˆxN)2]=u(xα)−α(xαN−ˆxN)2,xα∈¯B1(ˆx). |
Then
α(xαN−ˆxN)2≤u(xα)−u(ˆx)≤2, |
and
limα→∞xαN=ˆxN. | (2.4) |
Moreover, using the one dimensional symmetry, for any x∈¯B1(ˆx) we have
u(x)−α(xN−ˆxN)2≤u(xα)−α(xαN−ˆxN)2=u(ˆx1,…,ˆxN−1,xαN)−α(xαN−ˆxN)2. |
Hence u(x)−α(xN−ˆxN)2 has a maximum in (ˆx1,…,ˆxN−1,xαN)∈B1(ˆx) for α large in view of (2.4). Then
u3(ˆx1,…,ˆxN−1,xαN)−u(ˆx1,…,ˆxN−1,xαN)≤P−k(diag(0,…,0,2α))=0. |
We deduce that u(ˆx1,…,ˆxN−1,xαN)≥0 for every α big enough. Using semicontinuity and (2.4) we conclude
u(ˆx)≥lim supα→∞u(ˆx1,…,ˆxN−1,xαN)≥0. |
ii) By contradiction, let us assume that u(x)=v(xN) is a classical solution of (2.1) and that v(t0)≠0 for some t0∈R. By i), v(t0)>0. Let
δ−=inf{t<t0:v>0in[t,t0]}andδ+=sup{t>t0:v>0in[t0,t]}. |
If δ−=−∞ and δ+=+∞, then v>0 in R and, since v3−v<0, we deduce by the Eq (2.1) that v″ for any t\in \mathbb{R} . In particular v is concave in \mathbb R , a contradiction to v > 0 .
If \delta^- > -\infty , then v(t) > 0 for any t\in(\delta^-, t_0] and v(\delta^-) = 0 . Moreover there exists \xi\in(\delta^-, t_0) such that v'(\xi) > 0 . Using the Eq (2.1) we deduce that v''\leq0 in [\delta^-, t_0] , hence v'(\delta)\geq v'(\xi) > 0 . This implies that for \varepsilon small enough v(\delta^–\varepsilon) < 0 , a contradiction to i).
The case \delta^+ < +\infty is analogous.
iii) Let u(x) = \tanh\left(\frac{x_N}{\sqrt{2}}\right) . Then
D^2u = {\rm diag}(0, \ldots, 0, u^3-u). |
If x_N\geq0 then u^3-u\leq0 and \mathcal{P}^-_{k}(D^2u) = u^3-u , while if x_N < 0 the function u fails to be a solution since \mathcal{P}^-_{k}(D^2u) = 0 < u^3-u .
Instead we claim that
\tilde u(x) = \begin{cases} \tanh\left(\frac{x_N}{\sqrt{2}}\right) & \text{if $x_N\geq0$}\\ 0 & \text{otherwise} \end{cases} |
is a viscosity solution of (2.1). This is obvious for x_N\neq0 . Now take \hat x = (\hat x_1, \ldots, \hat x_{N-1}, 0) . Since there are no test functions \varphi touching \tilde u by above at \hat x , automatically \tilde u is a subsolution. Let us prove the \tilde u is also a supersolution. Let \varphi\in C^2(\mathbb{R}^N) such that \varphi(\hat x) = \tilde u(\hat x) = 0 and \varphi\leq u in B_{\delta}(\hat x) . Our aim is to show that \lambda_{N-1}\left(D^2\varphi(\hat x)\right)\leq0 , from which the conclusion follows.
Let W_0 = \left\{w\in\mathbb R^N\, :\; w_N = 0\right\} and for any w\in W_0 such that |w| = 1 let
g_w(t) = \varphi(\hat x+tw) \qquad t\in(-\delta, \delta). |
Since \varphi touches \tilde u from below at \hat x and \tilde u = 0 when x_N = 0 , we deduce that g_w(t) has a maximum point at t = 0 . Then
g_w''(0) = \langle D^2\varphi(\hat x)w, w\rangle\leq0. |
Using the Courant-Fischer formula
\begin{equation*} \begin{split} \lambda_{N-1}\left(D^2\varphi(\hat x)\right)& = \min\limits_{\dim W = N-1}\max\limits_{\;w\in W, \, |w| = 1}\langle D^2\varphi(\hat x)w, w\rangle\\ &\leq \max\limits_{w\in W_0, \, |w| = 1}\langle D^2\varphi(\hat x)w, w\rangle\leq0 \end{split} \end{equation*} |
as we wanted to show.
As above one can check that, for any c\geq0 , the one dimensional function
\tilde u(x) = \begin{cases} \tanh\left(\frac{x_N-c}{\sqrt{2}}\right) & \text{if $x_N\geq c$}\\ 0 & \text{if $|x_N| \lt c$}\\ -\tanh\left(\frac{x_N+c}{\sqrt{2}}\right) & \text{if $x_N\leq-c$, } \end{cases} |
see figure 1, is non monotone in the x_N -direction and it is a solution of (2.1).
iv) By contradiction suppose that u(x) = v(x_N) is a positive viscosity supersolution of (2.1). We claim that v is strictly concave, leading to a contradiction with v > 0 in \mathbb R .
We first prove that v satisfies the inequality
\begin{equation} v''(t) \lt 0 \quad\text{for any $t\in \mathbb{R}$} \end{equation} | (2.5) |
in the viscosity sense. For this let \varphi\in C^2(\mathbb{R}) be test function touching v from below at t_0 . If we consider \varphi as a function of N variables just by setting \tilde\varphi(x) = \varphi(x_N) , then \tilde\varphi is a test function touching u from below at (x_1, \ldots, x_{N-1}, t_0) , for any (x_1, \ldots, x_{N-1})\in\mathbb R^{N-1} . Hence
\mathcal{P}^-_{k}\left({\rm diag} (0, \ldots, 0, \varphi''(t_0))\right)\leq\varphi^3(t_0)-\varphi(t_0) \lt 0 |
and then necessarily \varphi''(t_0) < 0 .
If v was not strictly concave, then there would exist t_1 < \bar t < t_2\in \mathbb{R} such that
\min\limits_{t\in[t_1, t_2]}\left(v(t)-v(t_1)-\frac{v(t_2)-v(t_1)}{t_2-t_1}(t-t_1)\right) = v(\bar t)-v(t_1)-\frac{v(t_2)-v(t_1)}{t_2-t_1}(\bar t-t_1) |
Then using \varphi(t) = v(t_1)+\frac{v(t_2)-v(t_1)}{t_2-t_1}(t-t_1) as a test function in (2.5) we obtain a contradiction.
v) By iv) there exists t_0\in\mathbb R such that v(t_0) = 0 . By the monotonicity assumption we get v(t) = 0 for any t\leq t_0 .
This section is concerned with the existence of entire radial solutions of the equation
\begin{equation} \mathcal{P}^-_{k}(D^2u)+f(u) = 0 \quad \text{in $\mathbb R^N$}, \end{equation} | (3.1) |
where f:\mathbb R\mapsto\mathbb R satisfies the following assumptions: there exists \delta > 0 such that
\begin{equation} \begin{cases} f\in C^1((-\delta, \delta))\;\;\text{and it is nondecreasing in $(-\delta, \delta)$}\\ f(u) \gt 0\;\;\;\forall u\in(0, \delta)\\ f(0) = 0. \end{cases} \end{equation} | (3.2) |
Prototypes of such nonlinearities are
f(u) = \alpha u+\beta |u|^{\gamma-1}u |
with \alpha > 0 , \gamma > 1 and any \beta\in\mathbb R .
Proposition 2. Under the assumptions (3.2) there exist infinitely many positive and bounded radial (classical) solutions of the Eq (3.1).
Proof. For any \alpha\in[0, \delta) let v_\alpha be the solution of the initial value problem
\begin{equation} \begin{cases} v_\alpha'(r)+\frac rk f(v_\alpha(r)) = 0, \;\;r\geq0\\ v_\alpha(0) = \alpha\, \end{cases} \end{equation} | (3.3) |
defined in its maximal interval I_\alpha = [0, \rho_\alpha) . Since v_0\equiv0 , then v_\alpha(r) > 0 for any r\in I_\alpha if \alpha > 0 . We claim that I_\alpha = [0, \infty) . For this first note that v'_\alpha(r) is nonpositive in a neighborhood of the origin since, using (3.2)–(3.3), one has
v'_\alpha(0) = 0\qquad\text{and}\qquad v''_\alpha(0) = -\frac 1kf(\alpha) \lt 0. |
If there was \xi_\alpha\in(0, \rho_\alpha) such that v'_\alpha(r) < 0 in (0, \xi_\alpha) and v'_\alpha(\xi_\alpha) = 0 , then by monotonicity 0 < v_\alpha(\xi_\alpha) < \alpha and by (3.3) we should obtain that f(v_\alpha(\xi)) = 0 . But this is in contradiction with (3.2). Hence v_\alpha is monotone decreasing and positive, so I_\alpha = [0, \infty) . Using again (3.3) we deduce moreover that \lim_{t\to\infty}v_\alpha(r) = 0 and for any r > 0
\begin{equation} \begin{split} v''_\alpha(r)& = \frac{v'_\alpha(r)}{r}-\frac rk f'(v_\alpha(r))v'_\alpha(r)\\ &\geq\frac{v'_\alpha(r)}{r}\, . \end{split} \end{equation} | (3.4) |
In the last inequality we have used the facts that v_\alpha is monotone decreasing, 0 < v_\alpha(r) < \alpha for any r > 0 and that f is nondecreasing in [0, \delta) by assumption.
By (3.3)
-\frac{v_\alpha^\prime(r)}{f(v_\alpha(r))} = \frac{r}{k} |
and a straightforward computation gives that the function v_\alpha can be written as
\begin{equation} v_\alpha(r) = F^{-1}\left(\frac{r^2}{2k}\right), \end{equation} | (3.5) |
F^{-1} being the inverse function of F(\tau) = \int_\tau^\alpha\frac{1}{f(s)}\, ds in (0, \alpha] .
From the above we easily deduce that for any \alpha\in(0, \delta) the radial function
\begin{equation} u_\alpha(x) = v_\alpha(|x|) \end{equation} | (3.6) |
is a positive radial solution of (3.1).
In the model case f(u) = u-u^3 the assumptions (3.2) are satisfied with \delta = \frac{1}{\sqrt{3}} . Moreover, for \alpha\in(0, \frac{1}{\sqrt{3}}] , we have
F(\tau) = \int_\tau^\alpha\frac{1}{s-s^3}ds = \log\frac{\alpha}{\sqrt{1-\alpha^2}}-\log\frac{\tau}{\sqrt{1-\tau^2}}\, , \quad\text{for}\;\tau\in(0, \alpha] |
and
F^{-1}\left(\frac{r^2}{2k}\right) = \frac{1}{\sqrt{1+e^{\frac{r^2}{k}+\log\frac{1-\alpha^2}{\alpha^2}}}}\, . |
Hence, by (3.5)–(3.6), we infer that the functions
u_\alpha(x) = \frac{1}{\sqrt{1+e^{\frac{|x|^2}{k}+\log\frac{1-\alpha^2}{\alpha^2}}}}\, , |
see figure 2, are smooth and positive radial solutions of
\mathcal{P}^-_{k}(D^2u)+u-u^3 = 0 \quad \text{in $\mathbb R^N$}. |
Let us explicitly remark that the condition \alpha\in(0, \frac{1}{\sqrt{3}}] ensures the validity of the inequality (3.4).
Let us recall that in [9], given a domain \Omega , following Berestycki, Nirenberg and Varadhan [8] we define
\mu_k^-(\Omega) = \sup\{\mu\in\mathbb R: \exists \phi\in USC(\Omega), \;\phi \lt 0\ \mbox{ s.t. }\ \mathcal{P}^-_{k}(D^2\phi)+\mu\phi\geq 0\ \mbox{ in } \Omega\}. |
We proved many results among them that \mu_1^- could be called an eigenvalue since under the right conditions on \Omega we construct \psi < 0 solution of
\mathcal{P}^-_{1}(D^2\psi)+\mu_1^-\psi = 0\ \mbox{in}\ \Omega, \ \psi = 0\ \mbox{on}\ \partial\Omega. |
But, differently from the uniformly elliptic case, \mu_1^- does not satisfy the Faber-Krahn inequality, see [11].
Another feature of \mu_k^- is that it is the upper bound for the validity of the maximum principle. Precisely if \mu < \mu_k^- and
\mathcal{P}^-_{k}(D^2v)+\mu v \leq 0\ \mbox{in}\ \Omega, \ v\geq 0 \ \mbox{on}\ \partial\Omega |
then v\geq 0 in \Omega .
It is well known that for uniformly elliptic operators, the principal eigenvalue goes to infinity when the domain decreases to a domain with zero Lebesgue measure.
In this section we shall construct a sequence of domains Q_n\subset\mathbb R^2 that collapse to a segment such that \mu_1^-(Q_n) the principal eigenvalue of \mathcal{P}^-_{1} stays bounded above by 1. Hence not only they collapse to a zero measure set, but they are narrower and narrower.
Indeed we consider
Q_n = \{(x, y)\in \mathbb{R}^2, 0 \lt \frac{nx+y}{2} \lt \pi, \ \frac{-\pi}{2} \lt \frac{nx-y}{2} \lt \frac{\pi}{2}\}. |
And we define
w_n(x, y) = -(\sin(nx)+\sin y) = -2\sin\left(\frac{nx+y}{2}\right)\cos\left(\frac{nx-y}{2}\right). |
Obviously w_n < 0 in Q_n and w_n = 0 on \partial\Omega . We shall prove that
\begin{equation} \mathcal{P}^-_{1}(D^2w_n)+w_n\leq 0 \ \mbox{ in }\ \Omega. \end{equation} | (4.1) |
This will imply that \mu_1^-(Q_n)\leq 1 . Indeed if 1 < \mu_1^-(Q_n) the maximum principle would imply that w_n\geq 0 which is a contradiction.
Let us show (4.1). Clearly
D^2w_n = \left(\begin{array}{cc} n^2\sin(nx) & 0\\ 0 & \sin y \end{array}\right). |
We shall divide Q_n in three areas. In Q_n\cap(0, \frac{\pi}{n})\times (0, \pi) both \sin(nx) and \sin y are positive, then
\mathcal{P}^-_{1}(D^2w_n) = \min(n^2\sin(nx), \sin y)\leq \sin y\leq \sin(nx)+\sin y = -w_n. |
In Q_n\cap (\{-\frac{\pi}{2n} < x\leq 0\}\cup \{\frac{\pi}{n}\leq x < \frac{3\pi}{2n}\}) where \sin(nx)\leq 0 and \sin y\geq 0 , then
\mathcal{P}^-_{1}(D^2w_n) = n^2\sin(nx)\leq\sin(nx)\leq \sin(nx)+\sin y = -w_n. |
Finally, in Q_n\cap (\{\pi\leq y < \frac{3\pi}{2}\}\cup \{-\frac{\pi}{2} < y\leq0\}) where \sin(nx)\geq 0 and \sin y\leq 0 we have
\mathcal{P}^-_{1}(D^2w_n) = \sin(y)\leq \sin(nx)+\sin y = -w_n. |
This ends the proof.
Remark 2. During the revision of the present paper a new preprint, [22], which improves the result of this Section, has been uploaded on ArXiv. In particular, applying [22, Corollary 3.2] to Q_n , one would obtain that for any n\in\mathbb N , \mu_1^-(Q_n) = \frac14 .
This research is partially supported by INDAM-GNAMPA.
The authors declare no conflict of interest.
[1] |
Aronson DG, Weinberger HF (1978) Multidimensional nonlinear diffusion arising in population genetics. Adv Math 30: 33-76. doi: 10.1016/0001-8708(78)90130-5
![]() |
[2] |
Ambrosio L, Cabré X (2000) Entire solutions of semilinear elliptic equations in \mathbb{R}^3 and a conjecture of De Giorgi. J Am Math Soc 13: 725-739. doi: 10.1090/S0894-0347-00-00345-3
![]() |
[3] |
Barlow MT, Bass RF, Gui C (2000) The Liouville property and a conjecture of De Giorgi. Commun Pure Appl Math 53: 1007-1038. doi: 10.1002/1097-0312(200008)53:8<1007::AID-CPA3>3.0.CO;2-U
![]() |
[4] |
Berestycki H, Caffarelli LA, Nirenberg L (1997) Monotonicity for elliptic equations in unbounded Lipschitz domains. Commun Pure Appl Math 50: 1089-1111. doi: 10.1002/(SICI)1097-0312(199711)50:11<1089::AID-CPA2>3.0.CO;2-6
![]() |
[5] |
Berestycki H, Hamel F, Monneau R (2000) One-dimensional symmetry of bounded entire solutions of some elliptic equations. Duke Math J 103: 375-396. doi: 10.1215/S0012-7094-00-10331-6
![]() |
[6] | Berestycki H, Hamel F, Nadirashvili N (2005) The speed of propagation for KPP type problems. I: Periodic framework. J Eur Math Soc 7: 173-213. |
[7] |
Berestycki H, Nirenberg L (1991) On the method of moving planes and the sliding method. Bol Soc Bras Mat 22: 1-37. doi: 10.1007/BF01244896
![]() |
[8] |
Berestycki H, Nirenberg L, Varadhan S (1994) The principle eigenvalue and maximum principle for second order elliptic operators in general domains. Commun Pure Appl Math 47: 47-92. doi: 10.1002/cpa.3160470105
![]() |
[9] | Birindelli I, Galise G, Ishii H (2018) A family of degenerate elliptic operators: Maximum principle and its consequences. Ann Inst H Poincaré Anal Non Linéaire 35: 417-441. |
[10] | Birindelli I, Galise G, Ishii H (2020) Existence through convexity for the truncated Laplacians. DOI:10.1007/s00208-019-01953-x. |
[11] | Birindelli I, Galise G, Ishii H (2020) Towards a reversed Faber-Krahn inequality for the truncated Laplacian. Rev Mat Iberoam 3: 723-740. |
[12] | Birindelli I, Galise G, Ishii H (2020) Positivity sets of supersolutions of degenerate elliptic equations and the strong maximum principle. arXiv:1911.08204. |
[13] |
Birindelli I, Galise G, Leoni F (2017) Liouville theorems for a family of very degenerate elliptic nonlinear operators. Nonlinear Anal 161: 198-211. doi: 10.1016/j.na.2017.06.002
![]() |
[14] |
Birindelli I, Lanconelli E (2003) A negative answer to a one-dimensional symmetry problem in the Heisenberg group. Calc Var Part Dif 18: 357-372. doi: 10.1007/s00526-003-0194-0
![]() |
[15] |
Birindelli I, Mazzeo R (2009) Symmetry for solutions of two-phase semilinear elliptic equations on hyperbolic space. Indiana Univ Math J 58: 2347-2368. doi: 10.1512/iumj.2009.58.3714
![]() |
[16] | Birindelli I, Prajapat J (2002) Monotonicity and symmetry results for degenerate elliptic equations on nilpotent Lie groups. Pacific J Math 201: 1-17. |
[17] |
Del Pino M, Kowalczyk M, Wei J (2011) On De Giorgi's conjecture in dimension N ≥ 9. Ann Math 174: 1485-1569. doi: 10.4007/annals.2011.174.3.3
![]() |
[18] | Farina A (1999) Symmetry for solutions of semilinear elliptic equations in \mathbb{R}^N and related conjectures. Ricerche Mat 48: 129-154. |
[19] | Farina A (2003) Rigidity and one-dimensional symmetry for semilinear elliptic equations in the whole of \mathbb{R}^N and in half spaces. Adv Math Sci Appl 13: 65-82. |
[20] |
Farina A, Valdinoci E (2011) Rigidity results for elliptic PDEs with uniform limits: An abstract framework with applications. Indiana Univ Math J 60: 121-141. doi: 10.1512/iumj.2011.60.4433
![]() |
[21] |
Ghoussoub N, Gui C (1998) On a conjecture of De Giorgi and some related problems. Math Ann 311: 481-491. doi: 10.1007/s002080050196
![]() |
[22] | Parini E, Rossi J, Salort A (2020) Reverse Faber-Krahn inequality for a truncated laplacian operator. arXiv:2003.12107. |
[23] |
Polacik P (2017) Propagating terraces in a proof of the Gibbons conjecture and related results. J Fixed Point Theory Appl 19: 113-128. doi: 10.1007/s11784-016-0343-7
![]() |
[24] | Savin O (2003) Phase transitions: Regularity of flat level sets, PhD Thesis of UT Austin. |
1. | Danping Chen, Animation VR scene mosaic modeling based on generalized Laplacian equation, 2022, 0, 2444-8656, 10.2478/amns.2022.2.0156 | |
2. | Matthieu Alfaro, Philippe Jouan, The Allen-Cahn equation with nonlinear truncated Laplacians: Description of radial solutions, 2024, 0, 1078-0947, 0, 10.3934/dcds.2024151 |