1. Introduction
Second-order ordinary differential equations (ODE) and systems are since Newton's second law of motion, one of the most study equations in mathematics and physics under many different situations. They also play a crucial role in the study of linear and nonlinear PDEs. In this work, we want to extend some results for second-order ODE found in the work of Mawhin, see [Reference Mawhin, Nussbaum, Fitzpatrick and Martelli23] to equations with the one-dimensional fractional laplacian instead of laplacian. In particular, we will prove the existence of periodic solutions to some basic classical models like Lienard [Reference Liénard20], Forbat and Huaux [Reference Huaux and Forbat15], Lazer–Solimini [Reference Lazer and Solimini18] for equations involving the fractional laplacian, see also other references in [Reference Mawhin, Nussbaum, Fitzpatrick and Martelli23].
Nonlocal operator with singular kernels, in particular, the fractional Laplacian has received much attention in recent years; this motived by many apply models in biology, physics, chemistry, finance, marine foreign, etc., where the underlining phenomena are governed by anomalous diffusion, connected with Levy flights where the fractional Laplacian appears naturally. As an example of applied phenomena where this type of operator appears we mention [Reference Buldyrev, Murphy, Prince, Viswanathan, Afanasyev and Eugene Stanley8, Reference Dyer, Humphries and Queiroz12, Reference Klafter and Metzler16, Reference Li and Zhang19, Reference Viswanathan, da Luz, Raposo and Stanley28] and a more mathematical review on the topic can be found in [Reference Bucur and Valdinoci7], see also the big list of reference in all these works. Notice that periodic patterns are naturally expected in many of these applied phenomena.
The fractional Laplacian operator can be defined for $s\in (0,1)$ via its multiplier $|\xi |^{2s}$ in Fourier space, notice $s=1$ corresponds to the Laplacian. It can also be defined by the formula
here $C_{N, s} > 0$ is a well-known normalizing constant and P.V stands for the principal value, see for example [Reference Bucur and Valdinoci7].
From the mathematical point of view, the equations involving the fractional Laplacian that we will study here required some new different technics since many tools from ODE such as energy method, integral factors and other key elements do not remain valid in the nonlocal case, and therefore we need different methods and arguments for the nonlocal case.
Before describing our main results, notice that the type of equations studied in [Reference Mawhin, Nussbaum, Fitzpatrick and Martelli23] are of the form
and
Mawhin was mainly interested in singular nonlinearities, that is, he supposes that $g$ becomes unbounded near the origin; the $+$ sign indicates that the particles have opposite charges, while the $-$ sign indicates that the particles have the same charge. So, it is said that the equation with the $+$ sign (resp. $-$) has an attractive singularity (repulsive resp.). We simply talk about the attractive and repulsive case.
Thus, we studied equation of the form
and
where $(\triangle )^{s}u(t):= - (-\triangle )^{s}u(t)$ and $s\in (1/2,1)$ from now on.
The main aim of the present paper, is to establish existence of periodic solutions to equations (1.4) and (1.5).
In the study of (1.4) we suppose that $f:(0, + \infty ) \to \mathbb {R}$, $g:(0, + \infty ) \to \mathbb {R}$ are $C^{\alpha }((0, + \infty ))$ and $e\in C^{\alpha }(\mathbb {R})$ with $\alpha \in (0,1)$. Without loss of generality, we assume that we are searching for $2\pi$-periodic solutions to simplify the discussion, therefore, we will assume that $e$ is also $2\pi$-periodic. Moreover, we will use denotation
its mean value.
Observe that the term with $u'$ is sometimes called drift term and corresponds to some ‘transport’ or ‘friction’ in some of the models.
Before giving our results, let us mention that periodic solutions are studied in [Reference del Mar González, DelaTorre, del Pino and Wei10, Reference Garcia-Melian, Barrios and Quaas14] without drift term and regular nonlinearity, see also [Reference Du and Gui11, Reference Navarro26, Reference Zhang, Gui and Du34].
Other types of periodic problems related to the spectral fractional Laplacian can be found in [Reference Ambrosio2–Reference Ambrosio and Bisci6].
Now, we will give our first existence result in the case of attractive singularity.
Theorem 1.1 Assume that the function $g:(0, + \infty ) \to \mathbb {R}$ is such that the following conditions hold:
(i) $g(t) \to + \infty$ as $t\to 0^{+}$.
(ii) $\limsup \limits _{t \to + \infty } g(t)<\bar {e}$.
Then equation (1.4) has at least one $2\pi$-periodic positive classical solution.
This theorem will be proved by using Perron's method. One of the main difficulty here is to find a periodic super-solution, this is obtained by solving a semi-linear problem of Lienard type. In fact, we have the following results for nonlocal Lienard-type equations.
Proposition 1.2 Let $f\in C^{\alpha }(0, + \infty )$ and $w\in C^{\alpha }_{2\pi }(\mathbb {R})$, then
has at least one $2\pi$-periodic classical solution $u$ if, only if $\bar {w}=0$. Moreover,
for some positive constant $M$.
To prove the necessary condition, we will use a basic property of fractional Laplacian for periodic functions, see lemma 2.2. To prove the existence we will use the Schauder-type estimates which are obtained by $H^{s}$ estimate that gives by the assumption $s\in (1/2, 1)$ a $C^{\alpha }$ estimate then with the help of interpolation inequality in Hölder space we are able to manage the drift term, so we get our Schauder-type estimates.
This type of results can be generalized to the nonlocal Lienard vector equation inspired by [Reference Mawhin25]. More precisely, we will study vector equations of the form:
where
with $H: \mathbb {R}^{n} \to \mathbb {R}$ is $C^{2, \alpha }( \mathbb {R}^{n})$, $A$ is a $n\times n$- matrix and $e\in C^{\alpha }(\mathbb {R}, \mathbb {R}^{n})$ and $e$ is also $2\pi$-periodic.
Looking for existence results, we give a sufficient condition for (1.7) to have a $2\pi$-periodic solution.
Theorem 1.3 If $\bar {e}\in Im A$ and
then (1.7) has at least one $2\pi$-periodic classical solution.
Now we describe our second main results concerning equation (1.5) that is in the case of repulsive singular nonlinearity.
For that, we suppose that $c>0$, $e\in C^{\alpha }(\mathbb {R})$, $e$ is $2\pi$-periodic with $\bar {e} > 0$ and $g \in C^{\alpha }(0, + \infty )$ is a given function satisfying the following conditions:
(G1) $\limsup \limits _{t \to + \infty } [g(t) + \bar {e}]<0$,
(G2) $\lim \limits _{t \to 0^{+}} g(t) = + \infty$ and $g$ is monotone near zero,
(G3) There exists $\epsilon >0$ such that $g(\tau )^{2s-2-\epsilon }\int _{\tau }^{1} g(t)\,\textrm {d}t \to + \infty$ as $\tau \to 0^{+}$,
(G4) $g(t)\geq -at -b$ for some $a>0$ and $b\geq 0$ and all $t>0$.
Now we are in a position to give our second main theorem.
Theorem 1.4 Assume that the conditions $(G1)$, $(G2)$, $(G3)$ and $(G4)$ are fulfilled, then equation (1.5) has at least one $2\pi$-periodic positive classical solution.
Here, it is important to mention the main difficulties with respect to the local case ($s=1$) where a priori uniform bounds hold for $u$, $u'$. But in the nonlocal case we just found a $L^{2}$ bound for $u'$.
Besides that, the main idea to get the a priori bound is to prove a new general energy type identity for periodic solution (see lemma 4.1) that gives the formula for
inspired in [Reference García-Melián, Quaas, Barrios and Del Pezzo13] that give related formula in the half-line. Here $I(u,a,b)$ is the right side given in lemma 4.1. This identity together with a local regularity theory (Harnack inequality, and Schauder-type estimates) at the maximum point of a solution will help us to estimate $I(u,a,b)$, then using $(G3)$ we are able to find the a priori lower bound for our periodic solutions, for more details see $\S$ $4$. The rest of the proof is based on degree type arguments.
As a by product of theorem 1.4 and bifurcation from infinity, this again inspired from the works [Reference Mawhin, Nussbaum, Fitzpatrick and Martelli23, Reference Mawhin and Schmitt24], we find multiplicity results in a regime where a priori bounds are lost, see theorem 5.1, below.
Let us finish this introduction by mentioning that some reference of previous results for fractional equation with singular nonlinearities in a different situation from the studied here can be found in [Reference María, Barrios, De Bonis and Peral21, Reference Sciunzi, Squassina, Canino and Montoro29, Reference Zhang, Gui and Du34].
The paper is organized as follows. Section 2 is devoted to proving proposition 1.2 which is the semi-linear case of (1.4) and its generalization theorem 1.3. Sections 3 and 4 are devoted to establishing some results for the existence of at least one $2\pi$- periodic positive solution for (1.4) and (1.5) respectively. Finally, the solution multiplicity results will be presented in § 5.
2. The semi-linear periodic problems
In this section, first we will prove some preliminary results that will facilitate the proof of theorem 1.3 and proposition 1.2. Notice that by proposition 1.2 of [Reference Garcia-Melian, Barrios and Quaas14] the fractional Laplacian of a smooth, $2\pi$-periodic function reduces to
where
and
We consider the space $X$ defined in [Reference Garcia-Melian, Barrios and Quaas14] as the closure of the set of $2\pi$-periodic functions $u\in C^{1}(\mathbb {R})$ with the norm
where $X$ is a Hilbert space when provided with the inner product
The space $X$ possesses good embedding properties which follow directly from the trivial relation $\|u\|_{H^{s}(0,2\pi )} \leq \|u\|_{X}$ for $u\in X$.
Consider now the family of equations
We want to establish Schauder-type estimates for equation (2.3), but for this we need first a $C^{\alpha }$ estimate that will follow from a $H^{s}$ bound and the embedding in $C^{\alpha }$ by the fact that $s>1/2$. For that, let us write
which implies that $\bar {u}=0$ and $\bar {e}=0$. hence substituting this into (2.3) we get
where $C= \bar {v}$. With that change we have the following lemma.
Lemma 2.1 Let $u\in C^{2s + \alpha }$, for some $\alpha \in (0,1)$, be a classical solution of (2.4) with $\bar {e}=0$ and $\bar {u}=0$ then there exists $K>0$ such that
for some $K>0$.
Proof. Let $u$ is a $2\pi$-periodic solution of (2.4). Multiplying by $u$ and integrating the equation (2.4) we find
Using (2.1), together with the $2\pi$-periodicity of $u$ we find
Let $H\in C^{2}(\mathbb {R})$ such that
Then
Hence substituting in (2.6) and using Cauchy inequality we have
On the other hand, we claim that we have the following Poincaré-type inequality. There exist $K>0$ such that
Indeed, as $\bar {u}=0$ and using Cauchy inequality we get
as $|t - y|< 2\pi$ gives
where $K= (2\pi )^{2s}$. Hence (2.7) and (2.8) give
and as $H^{s}(0,2\pi )\hookrightarrow C^{\alpha }[0,2\pi ]$ for $s>\frac {1}{2}$ we get
The following result is a basic property of our interest, where we will need the constant $C(1, s)$ which is precisely given by
The proof can also be found in [Reference Navarro26].
Lemma 2.2 Let $u$ a $2\pi$-periodic function, then
Proof. Indeed, as $u$ is a periodic function, we can write its Fourier series:
Then,
so
where the second integral of the above is zero because we are integrating an odd function in a symmetric (with respect to zero) domain. Similarly to the previous computation, we have
Hence substituting this in (2.10) gives
Hence,
Proof of proposition 1.2 Assume that $u$ is a $2\pi$-periodic solution of (2.4), then integrating (2.4) from $0$ to $2\pi$
hence, using the $2\pi$- periodicity of $u$ and the notation of lemma 2.1 we have
which, together with lemma 2.2 implies $\bar {e}=0$.
For the existence, we will look at a priori estimates for $u\in C^{1,\alpha }$, with $\alpha \in (0,1)$. We first notice that $f(C +u) \in C^{\alpha }(\mathbb {R})$, indeed as $u'$ is bounded we get $u$ is Lipschitz so
Now, $f(C +u)u' \in C^{\alpha }(\mathbb {R})$, because $f$ and $u'$ are bounded, using (2.11) we get
Hence (2.12) gives
where $C=C(\|f\|_{C^{\alpha }(\mathbb {R})}, \alpha )$. As $s\in \left (\frac {1}{2},1\right )$ by ([Reference Silvestre30], proposition 2.9), we obtain that
where $C=C(\|f\|_{C(\mathbb {R})}, \alpha, s)$, and using the Interpolation inequalities ([Reference Krylov17],theorems 3.2.1) we obtain that, for any $\epsilon > 0$, there is a positive constant $C = C(\alpha, \|f\|_{C^{\alpha }(\mathbb {R})}, \epsilon, s)$, such that
Hence, substituting (2.15) in (2.14) and choosing $\epsilon = \frac {1}{2}$
Now, as $u$ is $2\pi$-periodic we have
so that (2.9) implies
substituting this into (2.16) we get
Then, as $s>\frac {1}{2}$ by ([Reference Silvestre30], proposition 2.8), we obtain that
where $C=C(\alpha,s)$ is a positive constant. As $e\in C^{\alpha }(\mathbb {R})$ together with (2.13) and (2.17) given
in this case $C= C(\alpha, \|f\|_{C^{\alpha }(\mathbb {R})}, s)$.
Now we complete the proof, if $z\in C^{\alpha }_{2\pi } (\mathbb {R})$, the equation
has a unique bounded classical solution $2\pi$-periodic solution $u\in C^{2s + \alpha }(\mathbb {R})$ (see lemma 3.1, [Reference Garcia-Melian, Barrios and Quaas14]). So we can define the map $K: C^{\alpha }_{2\pi } (\mathbb {R}) \to C^{2s +\alpha }_{2\pi } (\mathbb {R})$ by $K(z)= u$ where $u$ is the solution of
so that $K: C^{\alpha }_{2\pi } (\mathbb {R}) \to C^{1,\alpha }_{2\pi } (\mathbb {R})$ is compact because the injection of $C^{2s + \alpha }$ into $C^{1,\alpha }$ is compact since we have $s>1/2$. As a consequence, we defined $T: C^{1,\alpha }_{2\pi } (\mathbb {R}) \to C^{1,\alpha }_{2\pi } (\mathbb {R})$ by $T(z)=u$ where $u$ is solution of
so $T$ is compact and using the Schaeffer theorem, $T$ has at least one fixed point $u$ in $B[0,R]\subset C^{1,\alpha }_{2\pi } (\mathbb {R})$ so
which implies that
Consequently, $v= \bar {v} + u$ is a $2\pi$-periodic solution of (1.6).
Now we extend the previous results for systems of equations of the form
For simplicity to find a priori bounds of the system (2.20) only in the case $\lambda =1$.
Proof of theorem 1.3 First, we reduce the case a solutions with zero mean value, let us write
which implies that $\bar {v}=0$ and $\bar {e}=0$, then substituting this into (1.7) becomes
we get the equivalent system
As $\bar {e}\in Im A$, at least one $\bar {u}$ exists solving (2.22), so for each $\bar {u}$, we just have to find a $2\pi$-periodic solution of (2.23) $v\in C^{1, \alpha }(\mathbb {R}, \mathbb {R}^{n})$.
On the other hand, note that
where $F$ is the Hessian matrix of H, so substituting in (2.23) get
Now, we shall find a priori estimates, first will look estimates for $H^{s}((0, 2\pi ),\mathbb {R}^{n})$ where $s=\min \limits _{1\leq i\leq n} s_{i}$. We will use the usual norm, but it is important to mention that for $s \leq s_{i}$ we have
Now, assume $v$ is a $2\pi$-periodic solution of (2.24), multiplying for $v$ and integrating the equation
so that (2.26) together the $2\pi$-periodicity of $v$ gives
so
Hence substituting in (2.27) gives
By assumption we have
so that (2.31) using Hölder inequality implies
Therefore, the Poincaré-type inequality gives
and hence
Then, we shall find a priori estimates in Hölder spaces; as $s>\frac {1}{2}$ by ([Reference Silvestre30], proposition 2.8) gives
hence we get
where $C=C(n,s, \alpha, \|f\|_{C^{\alpha }(\mathbb {R})}, \|A\|)$, by ([Reference Silvestre30], proposition 2.9) we obtain
and using the Interpolation inequalities ([Reference Krylov17], theorems 3.2.1) together with the fact that $2s > 1$, we obtain that, for any $\epsilon > 0$, there is a positive constant $C = C(n,\alpha, \|f\|_{C^{\beta }(\mathbb {R})}, \|A\|, \epsilon, s)$, such that
Hence,
Thus,
Now, as $v$ is $2\pi$-periodic we have
so that (2.9) implies
Hence choosing $\epsilon = \dfrac {1}{4nC}$, we get
Now,
Hence substituting (2.40) and (2.41) into (2.35) gives
Substituting (2.42) in (2.34) we get
so
where $C=C(n, s, \alpha, \|f\|_{C^{\alpha }(\mathbb {R})}, \|A\|)$. Now, if $y_{i}\in C^{1, \alpha }_{2\pi }(\mathbb {R})$ the equation
has a unique bounded classical solution $2\pi$ -periodic solution $v_{i}\in C^{2s + \alpha }_{2\pi }(\mathbb {R})$, then $v(t)=(v_{1}(t), \cdots, v_{n}(t))$ exists such that it is a solution of
where $y(t)= (y_{1}(t),\cdots, y_{n}(t))$.
Finally, we define the operator $T: C^{1, \alpha }_{2\pi }(\mathbb {R}, \mathbb {R}^{n}) \to C^{1, \alpha }_{2\pi }(\mathbb {R}, \mathbb {R}^{n})$ by $T(y)=v$ where $v$ is the solution
this operator is compact, because the injection of $C^{2s + \alpha }_{2\pi }(\mathbb {R}, \mathbb {R}^{n})$ into $C^{1, \alpha }_{2\pi }(\mathbb {R}, \mathbb {R}^{n})$ is compact, so using the Schaeffer theorem, $T$ has at least one fixed point $v$ in $B(0,R) \subset C^{1, \alpha }_{2\pi }(\mathbb {R}, \mathbb {R}^{n})$, this is
which implies that
i.e. $v$ is a $2\pi$-periodic solution of (2.24). Consequently, $u= \bar {u} + v$ is a $2\pi$-periodic solution of (1.7).
3. Case of attractive singularity
The standard method of sub and super solutions provides the following existence theorem for the $2\pi$-periodic solutions of equation (1.4), the more difficult part is to construct super-solution and is the place where we use proposition 1.2.
Lemma 3.1 Assume that there exist $2\pi$- periodic $C^{2s + \alpha }$- functions $\eta$ and $\beta$ for some $\alpha \in (0,1)$ and $s\in (1/2,1)$ such that $\eta \leq \beta$ and
in $\mathbb {R}$. Then equation (1.4) has at least one $2\pi$-periodic classical solution $u\in C^{2s +\alpha }_{2\pi }(\mathbb {R})$ satisfying $\eta (x)\leq u(x) \leq \beta (x)$ for all $x\in \mathbb {R}$.
Proof. Let $u\in C^{1,\alpha }_{2\pi }(\mathbb {R})$, we define for each $x\in \mathbb {R}$
then $H\in C^{\alpha }_{2\pi }(\mathbb {R})$.
Thus we can define
and
where $v$ is the unique solution of
Then, we have $K \circ N: C^{1,\alpha }_{2\pi } (\mathbb {R})\to C^{1,\alpha }_{2\pi } (\mathbb {R})$ is continuous and compact, as proved in proposition (1.2). Notice that $N(C^{1,\alpha }_{2\pi } (\mathbb {R}))$ is bounded. Hence, by Schauder's fixed point theorem, $K\circ N$ has a fixed point $u$, i.e., $u$ is a solution of
Now we prove that $\eta (x)\leq u(x)\leq \beta (x)$ for all $x\in \mathbb {R}$, we first show that $u(x)\leq \beta (x)$. The other inequality is similar. We assume by contradiction that $\max (u(x) - \beta (x))= u(\bar t) - \beta (\bar t)> 0$, here $\bar t$ is the point where the maximum is attained. So we have $(\triangle )^{s}u(\bar t) - (\triangle )^{s}\beta (\bar t) \leq 0$ and also
this gives a contradict. Therefore, $u(x)\leq \beta (x)$.
Now we are in position to prove theorem 1.1, which is direct now.
Proof of theorem 1.1 By assumption 1, there exists a constant $\eta >0$ such that
and thus $\eta$ is a sub solution for (1.4) with $2\pi$-periodic. We now write $e(x)= \bar {e} + \tilde {e}$, then by assumption $2$, there exists $R>0$ such that $g(x) \leq \bar {e}$ for $x \geq R$ and by proposition 1.2, the equation
has one $2\pi$-periodic solution $v$, so we take $C$ sufficiently large such that $C + v(x)\geq \max (\eta, R)$ for all $x\in [0, 2\pi ]$. Hence we take $\beta (x)= C + v(x)$, gives
so that $\beta (x)\geq \eta$ is a super-solution for (1.4) with $2\pi$-periodic. Then using lemma 3.1 there exists a $2\pi$-periodic solution $u$ of (1.4) with $\eta \leq u(x) \leq \beta (x)$.
Corollary 3.2 Assume that the function $g:(0, + \infty ) \to (0, + \infty )$ is such that the following conditions hold.
(i) $g(x) \to + \infty$ as $x \to 0^{+}$,
(ii) $\limsup \limits _{x \to + \infty } g(x) = 0$.
Then, if $\bar {e}>0$ the equation (1.4) has a positive $2\pi$- periodic solution.
Proof. Follows directly from theorem 1.1.
Example 3.3 The Forbat-type equation fractional Laplacian that is
where $f\in C^{\alpha }(\mathbb {R})$ and $e\in C^{\alpha }(\mathbb {R})$ and $\overline {e -1}>0$. By the change of variable $u=v - C$ we have
so let us now take
the corollary 3.2 implies that equation (3.2) has at least one $2\pi$-periodic positive solution, i.e. the equation (3.1) has at least one $2\pi$-periodic solution $v$ such that $v (x)>C$ for all $x\in [0,2\pi ]$.
4. Case of repulsive singularity
In this section, we will prove the existence of a positive $2\pi$-periodic solution to (1.5). For that, let us consider the family of equations:
First, we want to find a priori bounds. For that the following lemma is very important and corresponds to a new energy type identity for the periodic solutions, whose proof is based on ideas of lemma 3.2 in [Reference García-Melián, Quaas, Barrios and Del Pezzo13].
Lemma 4.1 Let $u\in C^{1}_{2\pi }(\mathbb {R})$, then
for every $0< a< b$.
Proof. Fix $0< a< b$ and choose $\delta$ and $M$ with the restrictions $0<\delta < a$ and $M>b+\delta$. We consider the integral
where $A_{\delta,M}=([a,b]\times [-M,M]) \cap \{(x,y)\in \mathbb {R}^{2}:\ |y-x|\ge \delta \}$.
It is not hard to see that
We now split $A_{\delta,M}=A^{1} \cup A^{2} \cup A^{3} \cup A^{4}$, where
Since the region $A^{2}$ is the reflection of $A^{3}$, with respect to the line $y=x$ and the integrand in the last integral above is antisymmetric, we get
Now, using Green's formula, where the line integral is to be taken in the positive sense. Parameterizing the line integral we have
Now we can pass to the limit in $I_{\delta, M}$ as $M\to +\infty$ using dominated convergence and get
where $A_\delta ^{1}=([a, b]\times (b, +\infty ) \cap \{(x,y)\in \mathbb {R}^{2}: y \geq x + \delta \}$ and $A_\delta ^{2}=([a, b]\times (- \infty, a) \cap \{(x,y)\in \mathbb {R}^{2}: y \le x-\delta \}$.
Finally, will want to pass to the limit as $\delta \to 0$ in (4.5). Observe that, since $u\in C^{1}_{2\pi } (\mathbb {R})$, we have that $u'$ is bounded so $u$ is Lipschitz, hence for $y$ close to $b$ gives
and we also have for $y$ close to $a$
So the passing to the limit is justified in the first and second integral in the right-hand side of (4.5) by dominated convergence. As for the third integral, for being $u$ Lipschitz gives
so that,
as $\delta \to 0^{+}$. As for the double integral, we also have that
for $x$ and $y$ close. Therefore, we can pass to the limit in the right-hand side of (4.5).
Now, on the left-hand side of (4.5) using the regularity of $u$ and dominated convergence, it follows that we can pass the limit and get our identity.
To prove that $u$ has an upper bound, we will need the following lemma, which we will prove using Fourier series.
Lemma 4.2 Let $u$ a $2\pi$-periodic function, then
Proof. Now we write
so,
and
by orthogonality we have
Hence,
As mentioned in the introduction, the fact that we can't establish that $u'$ is uniformly bounded, we need to use local regularity theory, the following lemma is proved in [Reference Tan, Felmer and Quaas32] (see theorem 3.1).
Lemma 4.3 Let $s>\frac {1}{2}$, for $\gamma \in (0,1)$ we have $(\triangle )^{1 -s} : C^{2,\gamma }(\mathbb {R}) \to C^{2s +\gamma }(\mathbb {R})$ is continuous, i.e.
The following is local regularity result, we use ideas of [Reference Quaas, Chen and Felmer27], see also [Reference Silvestre30], we give it here for completeness. Moreover, the basic ideas are the key elements in propositions 2.8 and 2.9 of [Reference Silvestre30] that are also used in this paper.
Lemma 4.4 Let $u\in L^{\infty }(\mathbb {R})$ be a solution of
with $\delta >0$. Then there exist $\gamma >0$ and $C^{*}$ such that $u\in C^{2s + \gamma }_{loc}(\mathbb {R})$. Moreover,
Proof. Let $w$ be a solution of
where $\eta \in C^{\infty }(\mathbb {R})$ such that $\eta \equiv 0$ outside $(x_{0} - \delta, x_{0} + \delta )$ and $\eta \equiv 1$ in $[x_{0} - \frac {3\delta }{4}, x_{0} + \frac {3\delta }{4}]$. So we get
Then, since $(\triangle )^{s}((\triangle )^{1 - s}w)= w''$ we have
we can use theorem 1.1 of [Reference Silvestre and Caffarelli31], to obtain that there exists $\gamma$ such that
Hence, using lemma 4.3 and (4.7) we get
Now use again interpolation inequality to get a local regularity result with drift term.
Theorem 4.5 Let $u$ be a solution of
with $\delta >0$. Then there exists $\gamma >0$ and $\bar C>0$ such that
Proof. Now, by the lemma 4.4,
Using now the Interpolation inequalities ([Reference Krylov17], theorems 3.2.1) together with the fact that $2s > 1$, we obtain that, for any $\epsilon > 0$, there is a positive constant $C = C(\gamma, \epsilon, s)$, such that
Hence, choosing $\epsilon = \frac {1}{2C^{*}}$ we get out results from (4.9) and (4.8).
Lemma 4.6 [Interior Harnack inequality] Let $v$ be a classical solution of
and $v\geq 0$ in $\mathbb {R}$ with $f\in L^{\infty }(-1,1)\cap C(-1,1)$ and $\delta \in (0,1)$. Then there exists $C_{0}>0$ independent of $v$ and $\delta$ such that
For the proof we quote [Reference Topp, Dávila and Quaas33] where a much more general equations are considered, including zero order term and drift term. A parabolic version of Harnack inequality with a drift term can be found in [Reference Dávila and Chang-Lara9].
Lemma 4.7 Assume that $(G1)$ and $(G2)$ hold then there exist constants $R_{1} \!>\! R_{0} \!>\! 0$ such that for each possible $2\pi$-periodic solution $u$ of ( 4.1) there exist $t_{0}$, $t_{1} \in [0, 2\pi ]$ such that $u(t_{0}) > R_{0}$ and $u(t_{1}) < R_{1}$.
Proof. We shall assume that $u$ is a solution of (4.1) for some fixed $\lambda \in (0,1)$ then
But, for lemma 2.2 and using the $2\pi$-periodicity of $u'$ gives
We now notice that with assumption $(G2)$, we get that there exists $R_{0}>0$ such that
whenever $0< x\leq R_{0}$. Therefore, if $0< u(t)\leq R_{0}$ for all $t\in [0, 2\pi ]$, we obtain $g(u(t)) + \bar {e} >$ for those $t$ and hence
a contradiction to (4.10), thus there exists $t_{0}$ such that $u(t_{0}) > R_{0}$. On the other hand, assumption $(G1)$ implies the existence of some $R_{1} > R_{0}$ such that
whenever $u \geq R_{1}$. Then, if $u(t) \geq R_{1}$ for all $t\in [0,2\pi ]$, gives $g(u(t)) + \bar {e} < 0$ for those $t$ and
a contradiction to (4.10), thus there exists $t_{1}$ such that $u(t_{1}) < R_{1}$.
Lemma 4.8 Let $u$ a solution positive $2\pi$-periodic of (4.1) and assume $(G1)$ then there exists constant $R$ such that
Moreover we have that there exists $C>0$ such that
Proof. By lemma 4.7
Now, multiplying (4.1) for $u'$ and integrating the equation
since
together with lemma 4.2 and using Hölder inequality gives
so that,
which, together with (4.12) implies
therefore the result follows.
Lemma 4.9 Assume that $(G4)$ holds then for each $u$ a solution of (1.5) there exists $C>0$ (independent of $u$) such that
Proof. First we find a $L^{1}$ bound for $g$. Notice that $(G4)$ implies that
integrating over $[0,2\pi ]$
by lemma 4.8 and (4.10) we have
Secondly, we now consider the following problem
where $h\in L^{2}$. Let $w\in C^{2}(\mathbb {R})$ such that
Then, $w$ is a solution of
and since $(\triangle )^{s}((\triangle )^{1 - s}w)= w''$ we have
Using [Reference Ambrosio1, theorem 2.7] we have
and as we know that for $\mu =1 - s + \epsilon /2$ (see [Reference Tan, Felmer and Quaas32, theorem 3.1])
we obtain
Now, we want to bound $\|w\|_{C^{1,\mu }(-2\pi,4\pi )}$, it follows from (4.18) and (4.16) we need to estimate $[w']_{C^{\mu }}$. Let $x,y\in (-2\pi,4\pi )$ and using Hölder inequality we get
so that (4.20) implies,
Finally, let $u$ a solution of (1.5) by the above with $h(t)= cu'(t) + e(t)$ we have
Lemma 4.10 Assume $(G1)$,$(G2)$,$(G3)$ and $(G4)$ then there exists $r \in (0,R_{0})$ such that each $2\pi$-periodic solution of (4.1) satisfies $u(t) > r$ for all $t \in [0, 2\pi ]$.
Proof. Let $t^{n}_{3}$, $t^{n}_{4}$ be the minimum point and the maximum point of $u_n$ in $[0,2\pi ]$, notice that by lemma 4.7 we have $u_n(t^{n}_4)>R_0$.
Now we assume by contradiction that $u_n(t^{n}_3) \to 0$ as $n\to \infty$. First we suppose that $t^{n}_{4}< t^{n}_{3}$,
and multiplying (4.1) by $u'_n$, then
so that,
We need to bound the first term in the right-hand side, since by (4.11) the order terms in the right hand are bounded, to get a contradiction with $(G3)$ by the fact that $u_n(t^{n}_4)>R_0$ and $u_n(t^{n}_3) \to 0$.
Let us now define
Notice that for $n$ large $u_n(t_{3}^{n})<\frac {R_0}{4C_0}$ therefore $\delta _n$ is finite and
Now we claim that there exists $\delta _{0}>0$ such that $\delta _n>\delta _{0}$. Suppose the contrary, so there exists a sub-sequence (still denote by $n$) such that $\delta _{n} \to 0$. Define $v_{n}(t)=u_{n}(\delta _{n}t + t_{4}^{n})$, we have that
where $h:=g(u_{n}(\delta _{n}t + t_{4}^{n})) + e(\delta _{n}t + t_{4}^{n})\in L^{\infty }(-1,1)$ by the definition of $\delta _n$. By lemma 4.6,
as $\inf \limits _{(-1,1)} v_{n} = \dfrac {R_{0}}{4C_{0}}$, we have
Taking $\delta _{n}$ small, we obtain a contradiction and the claim follows.
Now we can use theorem 4.5, and with $\delta =\delta _0 >0$ independent of $u_n$ there exists $\bar C>0$ (independent of $n$) such that
with $\delta > 0$, this implies that
Now we simplify the notation and drop the $n$ index to estimate $\int \limits _{t^{n}_{4}}^{t^{n}_{3}} u_n'(x) (-\Delta )^{s} u_n(x)\,\textrm {d}x$.
By lemma 4.1 we have
Let $\rho >0$ we have
by lemma 4.9 together with $(G2)$,$(G4)$ and (4.15) we obtain
where $C^{*}= 4a\pi R + 4b\pi + \|u'\|_{L^{2}[-2\pi,4\pi ]} + \|e\|_{L^{2}[-2\pi,4\pi ]}$, and by (4.11) we have $C^{*}< + \infty$. Now, the second integral we know is bounded
Now from (4.24) we have
so that
To analyse the same integral when $y < t_{4}- \delta$ and $t_{4} < x< t_{3}$ we have
Hence, (4.25), (4.26), (4.27), (4.28) and (4.29) give that there exists $M_0>0$ independent of $n$ such that
Thus, as mentioned above, this inequality implies a contradiction and the result follows.
We now prove the following existence result for the $2\pi$-periodic solution of equation (1.5), we shall use the ideas of Continuation Theorem of [Reference Mawhin and Gains22, Reference Mawhin, Nussbaum, Fitzpatrick and Martelli23].
Proof of theorem 1.4. By the definition of $R_0$ and $R_1$ in lemma 4.7 we have
and
Using proposition 1.2 we can define a map $K:C^{\alpha }_{2\pi }(\mathbb {R}) \to C^{\alpha }_{2\pi }(\mathbb {R})$ by $K(z)= u$ where $u$ is a solution of
where $K$ is compact.
Now, let us define the map $N: C^{\alpha }_{2\pi }(\mathbb {R}) \to C^{\alpha }_{2\pi }(\mathbb {R})$ by
Define the continuous projectors $Q: C^{\alpha }_{2\pi }(\mathbb {R}) \to C^{\alpha }_{2\pi }(\mathbb {R})$ by the constant function
Let $\Omega = \{ u \in C^{2s +\alpha }_{2\pi }(\mathbb {R}): r < u(t) < R,\; t\in [0,2\pi ] \}$ define one parameter family of problems
Explicitly,
For $\lambda \in [0, 1]$, observe that we have by lemma 2.2
Therefore for all $\lambda \in (0, 1]$, problem (4.1) and problem (4.33) are equivalent. Hence, lemmas 4.8 and 4.10 imply (4.33) does not have a solution of $\partial \Omega \times (0,1]$. For example, $\lambda =0$ (4.33) is equivalent to the problem
then, applying $Q$ to both members of this equation and by lemma 2.2, we obtain
Multiplying by $u'$ the second of those equations, using lemma 4.2 and integrating we have
so that
this implies $u'=0$ hence $u$ is constant, and we know that the constant solutions of $QNa =0$ satisfy the inequality $r< a< R$. Thus we have proof that (4.33) has no solution on $\partial \Omega \times [0,1]$. Therefore, $deg(I- K((1-\lambda )QN + \lambda N), \Omega, 0)$ is well defined for all $\lambda \in [0,1]$ and by homotopy invariant of the degree we have
where $E\subset C^{2s +\alpha }_{2\pi }(\mathbb {R}):$ is the one-dimensional space of constant maps.
From here, we use (4.30) and (4.31) and basic degree properties to get $deg(I - KQN, \Omega \cap E, 0)\neq 0$. Thus, we can conclude equation (1.5) has at least one $2\pi$-periodic classical solution.
5. Bifurcation from infinity and multiplicity of the solutions
In this section, we discuss a multiplicity result. We find the existence of a continuum of positive solutions, bifurcating from infinity this together with our previous results will give a multiplicity of solutions. This result is based on ideas from [Reference Mawhin, Nussbaum, Fitzpatrick and Martelli23, Reference Mawhin and Schmitt24]. Here we will use the notation of the previous section.
The eigenvalue problem
has associated eigenvalue $\mu =0$ and the constant eigenfunction $u\equiv 1$. Conversely, periodic eigenfunctions associated with $\mu =0$ are necessarily constant (see proof the theorem 1.4), therefore $\mu =0$ is a simple eigenvalue.
We want to find positive $2\pi$-periodic solutions of the equation
We will assume that continuous functions $G: (0, + \infty ) \to [0, \infty )$ and $e\in C^{\alpha }_{2\pi }(\mathbb {R})$ satisfy the following conditions
(H1) $\lim _{t \to 0^{+}} G(t) = + \infty$,
(H2) $\lim _{t \to + \infty } G(t) = 0$,
(H3) $\bar {e} >0$,
(H4) $\int _{0}^{1} G(t)\,\textrm {d}t= + \infty$.
We have the following result.
Theorem 5.1 Assume that conditions $(H1)$, $(H2)$, $(H3)$ and $(H4)$ are satisfied. Then there exists $\eta > 0$ such that the following holds:
Proof. We now take $g(u)= G(u) - \mu u$, so (5.1) is of the form (1.5), and satisfying the following conditions:
(H1’) $\lim _{t \to 0^{+}} g(t) = + \infty$
(H2’) $\limsup _{t \to + \infty } [g(t) + \bar {e}]< 0$ for $\mu \geq 0$
(H3’) $\int _{0}^{1} g(t)\,\textrm {d}t= + \infty$.
Therefore the results of theorem 1.4 are valid for equation (5.1) when $\mu \geq 0$. Then, by the continuity of degree defined in theorem 1.4 there exists $\eta >0$ such that for $-\eta \leq \mu < 0$ that degree is not trivial. So, there exists $u$ a solution for (5.1) with $-\eta \leq \mu < 0$. Now, by $(H2)$ $N(u)=o(\|u\|)$ at $u= + \infty$ then, the fundamental theorem on bifurcation from infinity from a simple eigenvalue implies the existence of a continuum $\mathcal {C}_{\infty }$ of positive solution $(\mu, u)$ bifurcating from infinity at $\mu =0$, since the solutions for $\mu \geq 0$ are bounded, the bifurcation is for the left side.
Acknowledgments
A. Q. was partially supported by FONDECYT Grant # 1190282 and Programa Basal, CMM. U. de Chile. L. C. was partially supported by ANID # 21191475 and by ‘PIIC de la Dirección de Postgrado y Programas de la UTFSM.’