Constructions of Delaunay-type solutions for the spinorial Yamabe equation on spheres
Abstract
In this paper we construct singular solutions to the critical Dirac equation on spheres. More precisely, first we construct solutions admitting two points singularities that we call Delaunay-type solutions because of their similarities with the Delaunay solutions constructed for the singular Yamabe problem in [33, 36]. Then we construct another kind of singular solutions admitting a great circle as a singular set. These solutions are the building blocks for singular solutions on a general Spin manifold.
Keywords. Spinorial Yamabe; Singular Solutions; Delaunay-type Solutions.
1 Introduction and statement of the main result
Since the resolution of the Yamabe problem, much has been clarified about the behavior of solutions of the semilinear elliptic equation relating the scalar curvature functions of two conformally related metrics. One of the starting points for several recent developments was R. Schoen’s construction of complete metrics with constant positive scalar curvature on the sphere , conformal to the standard round metric, and with prescribed isolated singularities (see [37]). In analytical terms, it is equivalent to seeking for a function satisfying
(1.1) |
in the distributional sense with singular at every point of . Here we denote by the standard Riemannian metric on .
Eq. (1.1) and its counterpart on a general manifold are known as the singular Yamabe problem, and has been extensively studied. Just as the classical Yamabe problem in the compact setting, the questions concerning metrics of constant positive scalar curvature are considerably more involved. Remarkable breakthroughs and geometrically appealing examples were obtained by Schoen and Yau [38] and Schoen [37] when the ambient manifold is the -sphere . The former established that if admits a complete metric with scalar curvature bounded below by a positive constant, then the Hausdorff dimension of is at most , and the latter constructed several examples of domains that admit complete conformally flat metrics with constant positive scalar curvature, including the case where is any finite set with at least two points. Subsequently, Mazzeo and Smale [35] and Mazzeo and Pacard [33, 34] generalized the existence results, allowing to be a disjoint union of submanifolds with dimensions between and when the ambient manifold is a general compact manifold with constant nonnegative scalar curvature, and between and in the case .
In the past two decades, it has been realized that the conformal Laplacian, namely the operator appearing as the linear part of (1.1), falls into a particular family of operators. These operators are called conformally covariant elliptic operators of order and of bidegree , acting on manifolds of dimension . Many important geometric operators are in this class, for instance, the conformal Laplacian, the Paneitz operator, the Dirac operator, see also [11, 14, 21] for more examples. All such operators share several analytical properties, in particular, they are associated to the non-compact embedding of Sobolev space . And often, they have a central role in conformal geometry.
Let be an -dimensional spin manifold, , with a fixed Riemannian metric g and a fixed spin structure . The Dirac operator is defined in terms of a representation of the spin group which is compatible with Clifford multiplication. Let be the associated bundle, which we call the spinor bundle over . Then the Dirac operator acts on smooth sections of , i.e. , is a first order conformally covariant operator of bidegree . We point out here that the spinor bundle has complex dimension .
Analogously to the conformal Laplacian, where the scalar curvature is involved, the Dirac operator on a spin manifold has close relations with the mean curvature function associated to conformal immersions of the universal covering into Euclidean spaces. This theory is referred as the spinorial Weierstraß representation, and we refer to [3, 4, 26, 27, 28, 18, 32, 42, 43, 44] and references therein for more details in this direction. In a similar way as in the Yamabe problem, the spinorial analogue of the Yamabe equation (related with a normalized positive constant mean curvature) reads as
(1.2) |
where stands for the induced hermitian metric on fibers of the spinor bundle. One may also consider the equation with an opposite sign
(1.3) |
which corresponds to negative constant mean curvature surfaces. However, since the spectrum of is unbounded on both sides of and is symmetric about the origin on many manifolds (say, for instance ), the two problems (1.2) and (1.3) are of the same structure from analytical point of view.
Although conformally covariant operators share many properties, only few statements can be proven simultaneously for all of them. Particularly, the behavior of solutions of the conformally invariant equation (1.2) or (1.3) is still unclear. From the analytic perspective, some of the conformally covariant operators are bounded from below (e.g. the Yamabe and the Paneitz operator), whereas others are not (e.g. the Dirac operator). Some of them act on functions, while others on sections of vector bundles. For the Dirac operators, additional structure (e.g. spin structure) is used for defining it, and hence, more attention needs to be payed on such an exceptional case.
In this paper we initiate an investigation into the singular solutions of the nonlinear Dirac equation (1.2) when the ambient manifold is , which is perhaps the most geometrically appealing instance of this problem. As was described earlier, for a given closed subset , it is to find metrics which are complete on and such that satisfies Eq. (1.2) with . This is the singular spinorial Yamabe problem. Let us mention that, up until now, no existence examples have been known for the singular solutions of Eq. (1.2). Our first main result is follows:
Theorem 1.1.
Let be a pair of antipodal points, for , or an equatorial circle for . There is a one-parameter family of spinors solving the problem
(1.4) |
such that is a complete metric on . Moreover,
-
if is a pair of antipodal points, the family is parameterized by .
-
if is an equatorial circle, the family is parameterized by , where each is a bounded open set, for and is unbounded.
Remark 1.2.
Let us remark that Eq. (1.4), or more generally Eq. (1.2), is invariant under several Lie group actions. For instance, the canonical action of on spinors keeps the equation invariant (i.e. if is a solution of Eq. (1.4) then is also a solution, for every fixed ). Moreover, for the case , the spinor bundle has a quaternionic structure which commutes with Clifford multiplication, see for instance the construction in [19, Section 1.7] or [29, Page 33, Table III]. In these cases, Eq. (1.4) is invariant under the action of the unit quaternions on spinors. Therefore, in general, it is crucial to distinguish solutions of Dirac equations under various group actions. For instance, these symmetries were exploited in [30] to construct families of solutions on the sphere and the symmetry was used in [31] to exhibit also non-trivial solutions for the sub-critical Dirac equation. Thanks to our constructions, the solutions in the family obtained in Theorem 1.1 are distinguished via their parameterizations. And if is a group that keeps Eq. (1.4) invariant, our construction shows a larger family of singular solutions.
As we will see in Section 3, via a conformal change of the metric , problem (1.4) can be transformed to
(1.5) |
when is a pair of antipodal points and
(1.6) |
when is an equatorial circle, where . To obtain the results for Eq. (1.4) in consistence with similar results for the classical Yamabe equation, a fundamental idea is to express the equation (1.5) and (1.6) on the cylinder , or . By introducing the cylindrical coordinates :
for , one may be expecting that the ansatz
could turn Eq. (1.5) into a more manageable problem via a separation of variables process leading to a ”radial” solution . This is the very case for many elliptic problems (with a corresponding change of the exponent on ), including the Yamabe equation, fractional Yamabe equation [13] and the -curvature problem [25]. However, we point out that in the scalar case, there is a symmetrization process that behaves well with elliptic operators, reducing the problem to the study of an ODE. But when dealing with differential operators acting on vector bundles (spinor bundle in our case), one does not have a general symmetrization process. In particular, even on the Euclidean spaces , one cannot use the radial ansatz , for , to reduce a Dirac equation to an ODE system in terms of .
Notice that the spinorial Yamabe equation (1.5) (resp. (1.6)) contains (resp. ) unknown complex-functions, which is a considerably large number as grows. Instead of blindly “guessing” a particular ansatz, our starting point is the spin structure, or more precisely the spin representation. In fact, we use the matrix representation of Clifford multiplication to construct a “nice” function space for spinor fields which is invariant under the action of the Dirac operator , see in Section 2.3 for the definition. We find that the space is of particular interest from two perspectives (see Remark 2.1 below): First of all, when the dimension , encapsulates several important and special formulations of spinors which are of interest to particle physicists when they study quantum electrodynamic systems. Many important physical simulations have been obtained by using these special spinors, see for instance [15, 41, 46, 12]. The second perspective is that, spinors in reduce the equation (1.5) significantly in the sense that, for any dimension , Eq. (1.5) and (1.6) can be reduced to the following ODE systems of only two unknown functions
(1.7) |
and
(1.8) |
where . After using the Emden-Fowler change of variable and writing , in (1.7), we get a nondissipative Hamiltonian system of
(1.9) |
And, by writing and , we can transform (1.8) into
(1.10) |
which is a dissipative Hamiltonian system.
Let us denote by
the corresponding Hamiltonian energy for the systems (1.9). Notice that is constant along trajectories of (1.9). Moreover, the equilibrium points of are
(1.11) |
where is a saddle point and the other two are center points; then it follows easily that for there is a periodic solution of (1.9) at the level . We set for these periodic solutions, parameterized by their Hamiltonian energies. We distinguish a dichotomy within the set based on the sign of the Hamiltonian energy . Indeed, , where
We will call elements of , positive Delaunay-type solutions and elements of , sign-changing Delaunay-type solutions for Eq. (1.5). This terminology is based on the similarities between and the classical Delaunay solutions for the Yamabe problem. We will clarify more these similarities along the paper. Since any will not reach the rest point , we have is bounded away from for all . Besides the above existence results, we have the following bifurcation phenomenon for the solutions .
Theorem 1.3.
Let , the following facts hold for the system (1.9):
-
Let and such that . Then (1.9) has inequivalent solutions. Particularly, these solutions are given by the constant solution and periods of a solution with fundamental period .
-
The Hamiltonian energy as and is (locally) compact in the sense that converges in to the nontrivial homoclinic solution of (1.9). That is, there exists such that converges in to , where
Corollary 1.4.
Let , Eq. (1.5) has a one-parameter family of singular solutions on , parameterized by . Moreover, the following asymptotic estimates hold
-
•
,
-
•
as ,
-
•
as ,
for each . Moreover, if is the solution corresponding to , then there exists such that converges in to as , where is a constant spinor with and “” stands for the Clifford multiplication on spinors.
It is important here to notice the difference between the decay rate of singular solutions that we found in the previous Corollary and the one of regular solutions of (1.5), studied in [9]. Indeed, the decay rate of a regular solution is but the one of a singular solution is .
For the system (1.10) we have
Theorem 1.5.
Let , the system (1.10) with initial datum has a solution globally defined on . Moreover, there are exactly two types of initial data, which can be characterized by:
and
for , where
In particular,
-
is a bounded open set for all ;
-
if we set , then and ;
-
if we set , then and for some small ;
-
if , then as . To be more precise, we have
as ;
-
if , then is bounded from below by a positive constant for all and is unbounded as ; furthermore, up to a multiplication by constant, is upper bounded by for all large.
By setting , we call these unbounded solution the Delaunay-type solution for Eq. (1.6). As a direct consequence of Theorem 1.5, we have a characterization of singular solutions for Eq. (1.6) on .
Corollary 1.6.
Let , Eq. (1.5) has a one-parameter family of singular solutions on , parameterized by . Moreover, the following asymptotic estimates hold
and
for each
This paper is organized as follows. First, in Section 2, we lay down the necessary geometric preliminaries that we will need to formulate our problem, including the main ansatz that will be adopted to find our families of singular solutions. Next, in Section 3, we use the ansatz to formulate the problem as a Hamiltonian system in (autonomous in the case of a point singularity and non-autonomous in the case of a one dimensional singularity). In section 4, we study the properties of the solutions of the Hamiltonian system in the two cases. This allows us to prove Theorems 1.3 and 1.5.
2 Geometric preliminaries
2.1 General preliminaries about spin geometry
Let be an -dimensional Riemannian manifold (not necessarily compact) with a chosen orientation. Let be the set of positively oriented orthonormal frames on . This is a -principal bundle over . A spin structure on is a pair where is a -principal bundle over and is a map such that the diagram
commutes, where is the nontrivial double covering of . There is a topological condition for the existence of a spin structure, namely, the vanishing of the second Stiefel-Whitney class . Furthermore, if a spin structure exists, it need not be unique. For these results we refer to [19, 29].
In order to introduce the spinor bundle, we recall that the Clifford algebra is the associative -algebra with unit, generated by satisfying the relation for (here is the Euclidean scalar product on ). It turns out that has a smallest representation of dimension such that as -algebra. In case is even, this irreducible representations is uniquely determined, but it splits into non-equivalent sub-representations and as -representations. If is odd, there are two irreducible -representations and . Both of them coincide if considered as -representations.
Define the chirality operator with being a positively oriented orthonormal frame on . In case is even, we have act as on , and sections of (resp. ) are called positive (resp. negative) spinors. While if is odd, the chirality operator acts on as , . Hence, for odd, it will cause no confusion if we simply identify and as the same vector space, that is , and equip them with Clifford multiplication of opposite sign.
Associated to the above observations, the spinor bundle is then defined as
Note that the spinor bundle carries a natural Clifford multiplication, a natural hermitian metric and a metric connection induced from the Levi-Civita connection on (see [19, 29]), this bundle satisfies the axioms of Dirac bundle in the sense that
-
for any , and
-
for any and ,
where is the hermitian metric on ;
-
for any and ,
where is the metric connection on .
The Dirac operator is then defined on the spinor bundle as the composition
where denotes the Clifford multiplication .
Let us remark that there is an implicit g-dependence in the Clifford multiplication “” or “”. In fact, considering a simple case where we replace g with a conformal metric , the isometry from onto defines a principal bundle isomorphism lifting to the spin level. Then it induces a bundle isomorphism , , fiberwisely preserving the Hermitian inner product and sending to . In the sequel, when necessary, we shall write and , etc., to precise the underlying manifold and the metric g.
2.2 Spinor bundle and the Dirac operator on product manifolds
In this subsection our notation is close to [39]. Let be a product of Riemannian spin -manifolds , . We have
where is the Lie group homomorphism lifting the standard embedding .
The spinor bundle over can be identified with
That is, we always put the even dimensional factor in the place of . And the Clifford multiplication on can be explicitly given in terms of the Clifford multiplications on its factors. In fact, for , , and
we have
(2.1) |
where in case and odd we set and . Let us remark that there are different ways to formulate the Clifford multiplication (2.1), but such changes are equivalent. Indeed, due to the uniqueness of , any definition of the Clifford multiplication on can be identified with (2.1) via a vector bundle isomorphism (see the examples in the next subsection).
Let and be the Levi-Civita connections on and . By
we mean the tensor product connection on . Then, by (2.1), the Dirac operator on is given by
(2.2) |
where if both and are odd and if is even.
For the case even, we have the decomposition and, moreover, when restrict on those half-spinor spaces we get .
2.3 A particular ansatz in Euclidean spaces
Let be equipped with the Euclidean metric, then the spinor bundle is given by . Although, from the abstract setting, the Dirac operator can be given by
where is a orthonormal base of , we can have a more explicit representation of this operator. In fact the Dirac operator can be formulated as a constant coefficient differential operator of the form
(2.3) |
where is a linear map satisfying the relation
(2.4) |
for all .
Let us give a possible construction of these by using complex matrices with a block structure. We start with and the -dimensional Dirac operator , that is we have the pure imaginary unit. For is even, we define
where “Id” is understood to be the identity on . And, if is odd, we define
It is illuminating to consider this construction in low dimensions:
Example 1.
For , we have
Writing a spinor field in components as , we then have
(2.5) |
Thus, in this case, the Dirac operator is simply the Cauchy-Riemann operator.
Example 2.
Example 3.
For , we have
and
And for the product , we have . By considering a bundle isomorphism
for , one easily verifies the correspondence
which justifies (2.2). Meanwhile, for the product and the associated spinor bundle , we have the fiberwise isomorphism
for and such that the action of
on coincides with the action of on . This verifies (2.2). Note the analogy with dimension two.
We could continue this analysis. For general , one can compute the matrices , the chirality operator and, particularly when is even, the corresponding bundle isomorphism to decompose the Dirac operator in a product structure. However, these explicit formulas are seldom. It is always simpler to use the abstract setting of the Clifford module.
It is interesting to note that the aforementioned explicit formula for the Dirac operator motivates a “nice” function space which is invariant under the actions of the Dirac operator. More precisely, let us set
where stands for the complex unit sphere in the spin-module . Then, following the rule of the Clifford multiplication or the relation (2.4), it is easy to check that
Moreover, in order to make sure that is continuous at the origin, one may consider a further restriction to the subspace
Remark 2.1.
-
(1)
It is interesting to see that the specific ansatz provided in contains some important formulations of spinors, which are of interest to many physicists when they are dealing with spinor fields in quantum electrodynamics. In fact, to the best of our knowledge, it can be traced back to R. Finkelstein, R. LeLevier and M. Ruderman [15] in 1951 when they investigated a nonlinear Dirac equation in . By separating the time variable, the authors introduced a very special formulation of a spinor field, i.e.
(2.7) where is the spherical coordinates on . And subsequently, this ansatz has been commonly used in particle physics where spinors play a crucial role, see for instance [41, 46] and [12] for a -dimensional analogue. Now, in our setting, we understand that the above spinor field belongs to the sub-bundle . Consider the standard spherical coordinates
and
for , and , if we restrict to (i.e. the variable is separated out, treated as the time variable) and take
we soon derive that
which is exactly the latter one in (2.7).
-
(2)
Although the special ansatz (2.7) for a spinor has been known for over half a century, it is still new and important to have the family for general dimensions. Particularly, the ansatz in reduces the Dirac equation significantly. Indeed, for the semilinear equations of the form
(2.8) where is a given function, the ansatz in transforms it equivalently to
making the problem much easier to deal with.
- (3)
-
(4)
The space is somehow natural within spinor fields. Indeed, if one looks at the parallel spinors on and the Dirac bubbles [10] (corresponding to Killing spinors on the sphere), then one notices that they all belong to . Hence, we can think about as a generalized special class of spinors.
3 Set up of the problems
Let us consider the -sphere to be , where the coordinates is given by the standard stereographic projection from the north pole (here stands for the north pole). For clarity, we use the sub- or superscripts to indicate the underlying dimensions. By setting for the south pole, we can see that the manifold is conformally equivalent to . The conformal diffeomorphism can be explicitly formulated by
(3.1) |
where we have and .
This observation leads to some further considerations. Typical examples arise from the (connected) domain whose complement is an equatorial circle. Without loss of generality, we may consider the domain
Then we have the following conformal equivalence
(3.2) |
We now consider the solutions of the spinorial Yamabe equation on the sphere , that are singular at a prescribed closed set . More specifically, we will consider the problem
(3.3) |
when is given by a pair of antipodal points, say , or an equatorial circle .
Before discussing the Delaunay family of solutions to Eq. (3.3), let us recall the transformation formula of the Dirac operator under conformal changes (see [22, 24]):
Proposition 3.1.
Let and be two conformal metrics on a Riemannian spin -manifold . Then, there exists an isomorphism of vector bundles which is a fiberwise isometry such that
where and are the Dirac operators on with respect to the metrics and g, respectively.
In what follows, our discussions will be build upon this formula.
3.1 The singular set is a pair of antipodal points
In this setting, without loss of generality, we assume . Then, as a direct consequence of Proposition 3.1, we have that if is a solution to the equation
(3.4) |
then () is a solution to Eq. (3.3). Notice that since Eq. (3.4) has the same structure as (2.8), we shall look at solutions of the form
(3.5) |
Then, applying the Emden-Fowler change of variable and write and , we are led to consider the following system
(3.6) |
This system is easily integrated and is nondissipative, in particular, the Hamiltonian energy
is constant along solutions of (3.6).
The equilibrium points for system (3.6) are
And there is a special homoclinic orbit
(3.7) |
corresponding to the level set ; it limits on the origin as tends to , and encloses a bounded set in the first quadrant of the -plane, given by . It is easy to see that orbits not enclosed by this level set, i.e. those orbits in , must pass across the -axis and -axis. That is and must change sign. Observe that the equilibrium point is contained exactly in two orbits: the homoclinic one and the stationary orbit . Hence, for orbits in , we must have that for all . And thus, we have an unbounded one parameter family of periodic solutions
which induces correspondingly a family of singular solutions to Eq. (3.4) via (3.5). Remark that as and as for each . Therefore, these solutions give rise to distinguished singular solutions of Eq. (3.3).
3.2 The singular set is an equatorial circle
First of all, we need to observe that Eq. (3.3) can be interpreted as an equation on by a conformal change of the Riemannian metric on . Consider the product metric on , given in -coordinates by , where parameterizes the unit sphere . Then it follows from the conformal equivalence (3.2) that
And as a direct consequence of Proposition 3.1, we have that if is a solution to the equation
(3.8) |
then is a solution to Eq. (3.3) with being a bundle isomorphism.
Let us remark that the formula (2.2) on product manifolds indicates a way to construct singular solutions for Eq. (3.8). In fact, if is odd (hence ), then is even and we can consider a special spinor of the form so that Eq. (3.8) is reduced to
(3.9) |
where is a spinor on . And once again, by using the conformal formula in Proposition 3.1, Eq. (3.9) can be equivalently transformed to
(3.10) |
where for . And the solutions of (3.9) and (3.10) are in one-to-one correspondence via the identification for spinors.
Now, by considering the ansatz
and applying the change of variable , we can reduce Eq. (3.10) to the system
(3.11) |
where and .
If is even, then the spinor bundle on can be identified with and the Dirac operator can be formulated as
Hence, considering a spinor of the form for , we may reduce Eq. (3.8) to the following Dirac system
on . Similar to Eq. (3.10), we can transform the above system to
(3.12) |
on .
Now, using the ansatz
in and applying the change of variable , we then get the following system
(3.13) |
where we have substituted , , and . Therefore, we can consider the solutions for which and ; these are the solutions having the simplest and clearest structure. By writing and , we can turn (3.13) into
which exactly coincides with (3.11).
Clearly, the system (3.11) has an Hamiltonian structure, where the Hamiltonian energy is given by
It is evident that this system is dissipative and there is no periodic solution. However, one may consider solutions that are not converging to as . More precisely, we will characterize the following family of solutions
which induces a family of singular solutions to Eq. (3.9). Hence these solutions gives rise to singular solutions of Eq. (3.3). In this setting, we shall call the family the Delaunay-type solutions.
4 Analysis of the ODE systems
This section contains our main study of the dynamical systems (3.6) and (3.11). We point out that both systems have a variational structure. In fact, if we denote , systems (3.6) and (3.11) can be rewritten as
(4.1) |
where
and stands for the corresponding Hamiltonian energy. The functionals
and
can be used to obtain periodic solutions and homoclinic solutions for (4.1) respectively. In particular, there is one-to-one correspondence between -periodic solutions of (4.1) and critical points of (as long as is periodic in the -variable or independent of ). Similarly, critical points of correspond to homoclinic solutions of (4.1), i.e., as .
For the autonomous system, i.e. (3.6), we point out that the existence of a -periodic solution for every , some , and the asymptotic behavior of these solutions as have been already investigated in [45, 2]. By summarizing their results, we have
Proposition 4.1.
There exists such that for every the Hamiltonian system (3.6) has a non-constant -periodic solution . The family is compact in the following sense: for any sequence , up to a subsequence if necessary, converges in to a nontrivial solution of the system (3.6) on satisfying
i.e., is a homoclinic orbit.
Notice that the previous proposition does not provide a clear description of the behavior of the solutions as or a characterization of . For instance, from the arguments in [45, 2], we do not have an estimate of and we do not know if there are non-constant solutions below . In fact, if has a “good” structure around its equilibrium points, then one can use Lyapunov’s center theorem to exhibit a family of small amplitude periodic solutions bifurcating from the equilibrium solution and also have an estimate on . Nevertheless, this does not provide uniqueness of the family of non-constant solutions.
In the sequel, we will perform different approaches to characterize the Delaunay-type families and . We also want to point out that an alternative method can be used to find periodic solutions of family using variational analysis and by tracking the least energy solution, we can characterize the homoclinic energy , corresponding to the least energy solution for the functional . This procedure was used in a more general setting of product manifolds in [6].
4.1 The nondissipative case: Bifurcation of the positive periodic orbits
In order to analyse the dynamical system (3.6), we recall that
for and , which is independent of . We will focus on the periodic solutions/orbits of (3.6) in the first quadrant of the -plane, that is for all . Such solutions will be referred as positive solutions.
System (3.6) has an “obvious” constant solution for all . From now on, we intend to look at non-constant solutions. By setting and , we have and (3.6) becomes
(4.2) |
where we denote and for simplicity. After multiplication by in the second equation, we obtain
Thus, for any solution and , there exists a constant such that , that is,
(4.3) |
For , let us denote
Remark that, if is a non-constant -periodic solution of (4.2), then must achieve the maximum and minimum in one period. Hence has at least two zeros. This, together with the first equation in (4.2), implies that should vanish at least twice. Therefore, the conditions on are particularly restrictive. In fact, for , we can combine the first equation in (4.2) and (4.3) together to obtain . Then, if there exist and such that and , we have and . Clearly, this should corresponds to the homoclinic solution (3.7) and can not be periodic. For , by analyzing the algebraic equation , we can see that has exactly two zeros on given by the relations
But we find , which fails to satisfy the second inequality in (4.3). So the remaining range for is . However, it is obvious that can not be large.
Lemma 4.2.
If is small, has exactly two zeros on .
Proof.
We only prove the case , i.e. , since is much easier. Notice that
for and , we have and in for some small.
Observe that the two maps and have exactly two intersections for small enough. We denote the horizontal coordinates of these two intersections by . Then we have on and on . Therefore, is a strict local minimum, whereas is a strict local maximum.
Since (we used the facts , and ), we have for all small . Hence . This implies has exactly two zeros on . ∎
Let
We remark that, for , can not have a third zero in since changes sign at most twice and .
Lemma 4.3.
and has only one zero, which is the global maximum. Furthermore, for all and .
Proof.
Since is obvious, we only need to check the remaining statements. To begin with, we mention that
(4.4) |
provided that and . Hence, if for some and , we have for all . Moreover, due to the continuity of with respect to , there exists such that for . Therefore, we can see that is an open interval and that (otherwise will have two zeros). By choosing a sequence and such that , we have is bounded and as . Therefore has only one zero, which is the global maximum. The last assertion comes from the fact (4.4). ∎
Remark 4.4.
The value of can be explicitly computed. Precisely, we have
In fact, is the largest positive number such that the equation has a solution.
In the sequel, let , we set the points such that vanishes. It is worth pointing out that and are functions of . Then is positive on the interval . And Eq. (4.3) is now equivalent to
which can be solved by , where
and is a constant.
Of course, is defined on the interval . By noting that and are simple roots of (that is for ), we have is well-defined. Moreover, we have and as or . Therefore, has an inverse which increases from to on the interval . Now, solutions to (4.3) can be represented as for .
Setting
(4.5) |
it follows that is a -periodic solution of Eq. (4.2) and can not have smaller period. Moreover, this (jointly with the corresponding from Eq. (4.2)) gives rise to a positive solution of Eq. (3.6) with .
Lemma 4.5.
The mapping is continuous. Particularly,
Proof.
For starters, we shall write and to emphasize that and are functions of . Notice that and are solutions to the equation . By the implicit function theorem, we have and are functions, in particular,
Since we have assumed , we have
which implies that and .
The continuity of is obvious and, without digging out very much from the function , we can evaluate the asymptotic behavior of as goes to the end points and . In fact, to see the limiting behavior of as , we first observe that and for all small . Hence we have . Moreover since is the larger solution to the equation . Then
Thus, by taking , we have .
For close to , we set , that is
By writing and , we can write in its factorization
with
where
and
From elementary computations, we can simply write
(4.6) |
for . Then we can reformulate as
(4.7) |
Notice that, as approaches , we have . By the continuity of , we have
where
This completes the proof. ∎
Remark 4.6.
We end this section by comparing the classical Delaunay solutions that appear in the study of the singular Yamabe problem and the solutions that we have just studied above. Let us recall the classical Delaunay solutions for the singular Yamabe problem as in [33, 36], that are obtained by solving the ODE
(4.8) |
This equation is clearly nondissipative, and the corresponding Hamiltonian energy is
By examining the level sets of , we see that all bounded positive solutions of Eq. (4.8) lie in the region of the -plane where is non-positive. In the figures below, we show a few orbits for both the Hamitonians for the systems (3.6) and (4.8) when .


4.2 The dissipative case: Shooting method
In this subsection, we investigate the system (3.11). In particular, since we are looking for singular solutions of the spinorial Yamabe equation, we are interested in solutions of (3.11) such that
In order to avoid unnecessary complexity and to get non-trivial solutions, we choose as initial conditions
Moreover, the symmetry of the system allows us to consider only the case .
Lemma 4.7.
For any , there is , unique solution of (3.11) satisfying . Furthermore, depends continuously on , uniformly on , for any .
Proof.
To begin with, we may write the system (3.11) in integral form as
for . Since the right-hand side of the above equation is a Lipschitz continuous function of , the classical contraction mapping argument gives us a local existence of on . Let be the maximal interval of existence for .
Clearly, if we define and for , we have is a solution on . Suppose that . Then we have as .
Let us denote
A simple computation implies
so that the energy is non-increasing along the solution , on . However, since we have as , we find
as , which is absurd. Hence we have and are globally defined on . ∎
In what follows, we state some basic properties for solutions of (3.11).
Lemma 4.8.
Given , then the following holds:
-
•
If, for some , we have , then and .
-
•
If, for some , we have , then and .
Moreover, both and can not change sign infinitely many times in a bounded interval .
Proof.
Observe that the only rest point of system (3.11) is . Furthermore, for , the Cauchy problem for (3.11) is locally well-posed for any initial datum , for both and . Thus, a rest point cannot be reached in a finite time.
In order to see that both and can only change sign a finite number of times in a bounded interval , we assume by contradiction that there exists and in such that and as , for all , and (resp. ) changes sign a finite number of times on (resp. ) for any .
If , then will not change sign in a left neighborhood of and in a right neighborhood of . Then the first equation in (3.11) implies that has the same sign as , which is impossible. Hence . Similarly, one obtains . Therefore . Moreover, it can not happen that while (resp. ) keeps a definite sign around (resp. ). Therefore, we must have . In particular, we have , which is also impossible. ∎
Lemma 4.9.
Given . If is a bounded solution, i.e., for all and some , then as .
Proof.
By symmetry, we only need to prove the result for . Multiplying by (resp. ) the equations in (3.11), we have
Thus we need to show that as .
Suppose by contradiction that, for arbitrary small , there exists large such that
for some . Since
we find
Therefore, by enlarging , we can assume without loss of generality that . And hence, we obtain
which implies
By taking , we have as . This contradicts the boundedness of . ∎
Remark 4.10.
From the above result, we can conclude that, if there exists such that , the corresponding solution must be unbounded as (since the energy is decreasing).
Lemma 4.11.
Let . If is a solution such that . Then as .
Proof.
Since is decreasing, we can take such that and
for all Notice that can not reach in a finite time, we soon have
for all and depends only on . ∎
Lemma 4.12.
Let be a solution of (3.11) such that changes sign a finite number of times on , then there exists such that for all .
Proof.
Since changes sign a finite number of times on , we suppose without loss of generality that for all , some .
Assume, by contradiction, that for all . Then the second equation of (3.11) implies that for , that is, is increasing for . Hence we have
Notice that, by the second equation again, we have
We deduce that
Hence . However, since and have opposite sign, we find
as , which is impossible.
Let be such that . Then, it follow from the first equation of (3.11) that . If there exists such that and on , we soon derive that in a left neighborhood of . Thus, by the the first equation of (3.11) again, we get . This is impossible since we have assumed for all . Therefore, by taking , we conclude for all . ∎
Corollary 4.13.
Let be a solution of (3.11) such that changes sign a finite number of times on , then there exists such that for all .
Proof.
Suppose that there exists such that . Then and enters to negative values, and can not have further zeros. In fact, if there is such that and on . We will have , which is impossible. Then we obtain a contradiction with Lemma 4.12.
∎
Corollary 4.14.
Let be a bounded solution of (3.11) such that (or ) changes sign a finite number of times on , then
as .
Proof.
By virtue of Lemma 4.12 and Corollary 4.13, we can take large enough such that for all . Then, it can be derived from (3.11) that
Hence, from the boundedness of and , we have
(4.9) |
for sufficiently large, where is a constant.
Let and , for . One checks easily that
By taking such that
for some , we find
for all . Then, by the comparison principle, we have
for all , which completes the proof. ∎
Lemma 4.15.
Let be a solution of (3.11) such that changes sign a finite number of times on . If for all , then as , for some constants possibly depending on .
Proof.
We only prove the result for . Note that
where can be fixed arbitrarily small and the last inequality comes from Lemma 4.12. Therefore, we have
for all , which completes the proof. ∎
Now, for and the corresponding solution of (3.11), we introduce the sets , and defined for by
Notice that is a hyperbolic equilibrium point of the Hamiltonian energy for any . It is, then, immediate to see that as it includes the interval , since
As we will see later, tracking the sign changes of the solutions is crucial for the proof of Theorem 1.5. The main idea is to study the stratified structure of the solutions. This will be done by checking their topology and boundedness. The boundedness, allows us to track the of and allowing us to prove that all the sets are not empty. As we will see below, the idea of tracking the signs coming from a limiting problem with explicit solutions and infinitely many sign changes. This property will allow us to prove boundedness of the desired sets.
Let us start first by discarding the sets :
Lemma 4.16.
for all .
Proof.
Suppose to the contrary that for some . Let and be the corresponding solution. Then, by substituting into Eq. (3.11), we obtain
(4.10) |
This gives
for all . By Lemma 4.12, for large enough, we can divide the above inequality by to get
where we have assumed without loss of generality that and for large. Hence we have
(4.11) |
for some constant . And therefore, there exists such that for all . Now, by (4.10), we have
for , that is, is decreasing for all large .
Lemma 4.17.
There exists constants such that, if for some ,
-
;
-
;
-
changes sign times on ;
then .
Proof.
Suppose that , it remains to show that . Without loss of generality, let us assume that and . Set
If , we have changes sign at most once in . Indeed, as long as , the second equation of (3.11) implies that whenever vanishes. Therefore, can not change sign more than once. If does not change sign on , we have , which is absurd. However, if does change sign once in , we have for all large . This contradicts Lemma 4.12. Therefore, we have and .
Claim 1.
changes sign exactly once in .
In fact, by rewriting the second equation of (3.11), we have
for . If stays positive on , by Lemma 4.8, we have on a left neighborhood of , which is impossible.
To proceed, let us set and . Then satisfies the following system
(4.12) |
with Hamiltonian energy
Clearly, we have for . And, by Claim 1, we can make slightly larger so that on . That is, we have on , , and changes sign once in .
In what follows, we are going to prove that stays positive on . Then the second equation in (4.12) shows that for all . And hence for any . In this case, we have and for all , which implies for . That is, changes sign exactly once on . Therefore .
Suppose, by contradiction, that there exists such that and on . Then, the second equation in (4.12) implies that is decreasing on . And hence, . Then, we only need to consider the situation , since the condition will immediately trap the solution in the third quadrant of -plane for , and leads us to have .
In the case , by and , we have
Let be such that
and
By assuming suitably small, such and always exist, and we can have that . In fact, by setting
we have is nothing but the vanishing point of in the negative line, i.e.,
(4.13) |
and is the smallest point such that . Then, use the fact for all , we have
for . Hence, we deduce
(4.14) |
for . Notice that
and
By using the second equation in (4.12) and (4.14), we find
(4.15) | ||||
where depends only on (since we have assumed is small). On the other hand, we have
for , where in the last inequality we used and
Hence, by (4.15), we obtain
provided that and is small enough. This implies reaching a contradiction, and the proof is hereby completed. ∎
The next lemma provides the main properties of the sets and .
Lemma 4.18.
For all , we have
-
is an open set;
-
if , then there exists such that ;
-
if and is bounded, then ;
-
if both and are bounded, set , then there exists such that .
Proof.
is quite obvious, since it comes from the continuity of the solutions with respect to the initial datum.
To see , we fix . Then we have as . Given as in Lemma 4.17, there exists such that , and changes sign times on . The continuity of the solution with respect to implies that the same holds for an initial datum for small. Then the conclusion follows by Lemma 4.17.
To check , let us set and take a sequence such that as . If we suppose that for some , then suggests that for large. Hence we have . This implies which is absurd since is an open set. Notice that, by the continuity property of the solutions, the corresponding can change sign only a finite number of times on . Therefore we must have that for some . By , we have . This implies .
Finally, to see , we first observe that . Indeed, let be such that as , we have for any . This is because is an open set. Then, arguing similarly as in , we get that as claimed. Now, by , we have for some . Since we have assumed the boundedness of , we find . Thus . ∎
Our next result is the boundedness property of the sets and .
Proposition 4.19.
is bounded for each .
Before prove Proposition 4.19, let us do some preparations. Denoted by , we consider the following rescaling
We find the system for is
(4.16) |
together with the initial datum . The limiting problem associated to Eq. (4.16) is
(4.17) |
with .
Lemma 4.20.
Proof.
The Hamiltonian energy associated to (4.16) is given by
And it is easy to see that is decreasing along the flow, so that
This implies that
(4.20) |
for some constant independent of .
Proof of Proposition 4.19.
Suppose the contrary, that is unbounded for some . Then we can find a sequence such that as .
By taking , Lemma 4.20 implies that uniformly on as , for any fixed . Notice that the solution of Eq. (4.17) can be explicitly formulated:
We can take large enough so that changes sign times on . Then, by Lemma 4.20, we have changes times on for all large . However, due to and , we have should change sign only times on . And thus, we get a contradiction. ∎
Proof of Theorem 1.5.
Let . By Lemma 4.18, we have . Let now . Applying Proposition 4.19 and Lemma 4.18, we have for some . Thus . Let . We have ; and so, by Lemma 4.18, , and then and , for some . Iterating this argument, we construct two increasing sequences and , , with and , for some .
Next, we will show that as . Suppose, by contradiction, that is bounded and . We can see that for all . Indeed, if for some finite , it follows that will be trapped in one of the connected components of , for all . Since Lemma 4.8 implies that changes sign a finite number of times in , we have for some . This contradicts the definition of as is open. Moreover, must change sign infinite many times on .
Using the facts is decreasing on and bounded from below, we have . In particular,
(4.22) |
Multiplying by (resp. ) the equations in (3.11), we have
This implies
Hence we have , which shows that as . Since changes sign infinitely many times as , we have . This, together with (4.22), implies that as .
Therefore, one may take sufficiently large such that (where is given by Lemma 4.17), and changes sign times on . By Lemma 4.17, we have , reaching another contradiction.
Finally, in order to see that for , let us consider two possibilities: and . In the first case, we must have that as , which directly implies the assertion. In the latter case, we deduce that as . And hence converges to a positive constant. This shows that grows as for large.
Remark 4.21.
The numerical simulations performed on system (3.11) indicate the following. For each , starting from larger than some , the solution orbits will make a circle around a particular point (in either the first quadrant or the third quadrant) before going to infinity. As grows, the circle is becoming larger; and once the circle touches the origin, we will have a homoclinic solution of (3.11), which implies . The set seems to have only one point, and hence are just open intervals. In particular, we conjecture that is simply a countable set of discrete points. This is illustrated in the following Fig. 3, where numerical experiments are performed on a -dimensional system. The first row shows the solution orbits on with three different initial datum in , and specifically , and . The second and third rows show the solutions with initial datum and , respectively

Acknowledgements
Y.S. is partly supported by NSF grant DMS , ” Regularity vs singularity formation in elliptic and parabolic equations”.
References
- [1]
- [2] A. Abbondandolo, J. Molina, Index estimates for strongly indefinite functionals, periodic orbits and homoclinic solutions of first order Hamiltonian systems, Cal. Var. PDEs, 11 (2000), 395-430.
- [3] B. Ammann, A variational problem in conformal spin geometry, Habilitationsschrift, Universität Hamburg 2003.
- [4] B. Ammann, The smallest Dirac eigenvalue in a spin-conformal class and cmc immersions, Comm. Anal. Geom. 17 (2009), no. 3, 429-479.
- [5] C. Bär, Extrinsic bounds for eigenvalues of the Dirac operator, Ann. Global Anal. Geom. 16 (1998), no. 6, 573-596.
- [6] T. Bartsch, T. Xu, Curvature effect in the spinorial Yamabe problem on product manifolds. Cal. Var. PDEs. 61 (2022), no. 5, Paper No. 194, 35 pp.
- [7] W. Borrelli, Symmetric solutions for a 2D critical Dirac equation, accepted by Commun. Contemp. Math. 24 (2020), 2150019.
- [8] W. Borrelli, Stationary solutions for the 2D critical Dirac equation with Kerr nonlinearity, J. Diff. Equ. 263 (2017), pp. 7941-7964.
- [9] W. Borrelli and R. L. Frank, Sharp decay estimates for critical Dirac equations, Trans. Amer. Math. Soc. 373 (2020), pp. 2045-2070.
- [10] W. Borrelli, A. Malchiodi, R. Wu, Ground state dirac bubbles and Killing spinors, Comm. Math. Phys. 383 (2021), 1151-1180.
- [11] T. P. Branson, Second order conformal covariants, Proc. Amer. Math. Soc. 126 (1998), 1031-1042.
- [12] J. Cuevas–Maraver, P.G. Kevrekidis, A. Saxena, A. Comech, R. Lan, Stability of solitary waves and vortices in a 2D nonlinear Dirac model, Phys. Rev. Lett. 116:21 (2016), 214101.
- [13] A. DelaTorre, M. del Pino, M.d.M. Gonzalez, J.C. Wei, Delaunay-type singular solutions for the fractional Yamabe problem, Math. Ann. 369 (2017), 597-626.
- [14] H. D. Fegan, Conformally invariant first order differential operators, Quart. J. Math. Oxford, II. series 27 (1976), 371-378.
- [15] R. Finkelstein, R. LeLevier, M. Ruderman, Nonlinear Spinor Fields, Phys. Rev. 83, 326 (1951).
- [16] F. Finster, J. Smoller, and S.-T. Yau, Particlelike solutions of the einstein-dirac equations, Physical Review D, 59 (1999), p. 104020.
- [17] F. Finster, J. Smoller, S.T. Yau, Particle-like solutions of the Einstein-Dirac equations, Physical Review D. Particles and Fields. Third Series 59 (1999) 104020
- [18] T. Friedrich, On the spinor representation of surfaces in Euclidean 3-space, J. Geom. Phys. 28:1-2 (1998), 143-157.
- [19] T. Friedrich, Dirac Operators in Riemannian Geometry, Grad. Stud. Math., vol 25, Amer. Math. Soc., Providence (2000).
- [20] N. Ginoux: The Dirac Spectrum. Lecture Notes in Mathematics, vol. 1976. Springer, Berlin (2009).
- [21] R. Gover, L. J. Peterson, Conformally invariant powers of the Laplacian, Q-curvature, and tractor calculus, Comm. Math. Phys. 235 (2003), 339-378.
- [22] O. Hijazi, A conformal lower bound for the smallest eigenvalue of the Dirac operator and Killing spinors, Comm. Math. Phys. 104 (1986), 151-162.
- [23] O. Hijazi, Spectral properties of the Dirac operator and geometrical structures, in ”Geometric Methods for Quantum Field Theory”. Proceedings of the Summer School, eds. H. Ocampo et al.,Villa de Leyva, Colombia, July 12-30, 1999, World Scientific, Singapore, 2001, 116-169.
- [24] N. Hitchin, Harmonic spinors, Adv. Math. 14 (1974), 1-55.
- [25] A. Hyder, Y. Sire, Singular solutions for the constant -curvature problem, J. Funct. Anal. 280:3 (2021), 108819.
- [26] K. Kenmotsu, Weierstrass formula for surfaces of prescribed mean curvature, Math. Ann. 245:2 (1979), 89-99.
- [27] B. G. Konopelchenko, Induced surfaces and their integrable dynamics, Stud. Appl. Math. 96:1 (1996), 9-51.
- [28] R. Kusner, N. Schmitt, The spinor representation of surfaces in space, dg-ga/9610005.
- [29] H.B. Lawson, M.L. Michelson, Spin Geometry, Princeton University Press (1989).
- [30] A. Maalaoui, Infinitely many solutions for the spinorial Yamabe problem on the round sphere, NoDEA Nonlinear Differential Equations Appl. 23 (2016), pp. Art. 25, 14.
- [31] A. Maalaoui, Rabinowitz-Floer Homology for super-quadratic Dirac equations on spin manifolds. J. Fixed Point Theory Appl. 13 (2013), 175–199.
- [32] S. Matsutani, Immersion anomaly of Dirac operator on surface in , Rev. Math. Phys. 11:2 (1999), 171-186.
- [33] R. Mazzeo, F. Pacard, A construction of singular solutions for a semilinear elliptic equation using asymptotic analysis, J. Diff. Geom. 44 (1996), no. 2, 331-370.
- [34] R. Mazzeo, F. Pacard, Constant scalar curvature metrics with isolated singularities, Duke Math. J. 99 (1999), no. 3, 353-418.
- [35] R. Mazzeo, N. Smale, Conformally flat metrics of constant positive scalar curvature on subdomains of the sphere, J. Diff. Geom. 34 (1991), no. 3, 581-621.
- [36] R. Schoen, Variational theory for the total scalar curvature functional for Riemannian metrics and related topics, Topics in calculus of variations (Montecatini Terme, 1987), 120-154, Lecture Notes in Math., 1365, Springer, Berlin, 1989
- [37] R. Schoen, The existence of weak solutions with prescribed singular behavior for a conformally invariant scalar equation, Comm. Pure and Appl. Math. XLI (1988), 317-392.
- [38] R. Schoen, S. T. Yau, Conformally flat manifolds, Kleinian groups and scalar curvature, Invent. Math. 92 (1988) 47-71.
- [39] Y. Sire, T. Xu, A variational analysis of the spinorial Yamabe equation on product manifolds, Ann. Sc. Norm. Super. Pisa Cl. Sci. accepted.
- [40] S. Rota Nodari, Perturbation method for particle-like solutions of the Einstein-Dirac equations, Ann. Henri Poincaré 10 (2010), pp. 1377–1393.
- [41] M. Soler, Classical, stable, nonlinear spinor field with positive rest energy, Phys. Rev. D. 1 (1970), 2766-2769.
- [42] I. A. Taimanov, Modified Novikov-Veselov equation and differential geometry of surfaces, Amer. Math. Soc. Transl. (2) 179 (1997), 133-151; http://arxiv.org/dg-ga/9511005.
- [43] I. A. Taimanov, The Weierstrass representation of closed surfaces in , Funktsional. Anal. i Prilozhen. 32:4 (1998), 49-62; English transl., Funct. Anal. Appl. 32 (1998), 258-267.
- [44] I. A. Taimanov, The Weierstrass representation of spheres in , the Willmore numbers, and soliton spheres, Trudy Mat. Inst. Steklov. 225 (1999), 339-361; English transl., Proc. Steklov Inst. Math. 225 (1999), 225-243.
- [45] K. Tanaka, Homoclinic orbits in a first order superquadratic Hamiltonian system: convergence of subharmonic orbits, J. Diff. Equ. 94 (1991), 315-339.
- [46] M. Wakano, Intensely localized solutions of the classical Dirac-Maxwell field equations, Progr. Theoret. Phys. 35:6 (1966), 1117-1141.
Ali Maalaoui
Department of Mathematics,
Clark University,
Worcester, MA 01610-1477
[email protected]
Yannick Sire
Department of Mathematics, Johns Hopkins University,
3400 N. Charles Street, Baltimore,
Maryland 21218
[email protected]
Tian Xu
Center for Applied Mathematics, Tianjin University,
300072, Tianjin, China
[email protected]