Abstract

We study the stability of the cnoidal, dnoidal and snoidal elliptic functions as spatially-periodic standing wave solutions of the 1D cubic nonlinear Schrödinger equations. First, we give global variational characterizations of each of these periodic waves, which in particular provide alternate proofs of their orbital stability with respect to same-period perturbations, restricted to certain subspaces. Second, we prove the spectral stability of the cnoidal waves (in a certain parameter range) and snoidal waves against same-period perturbations, thus providing an alternate proof of this (known) fact, which does not rely on complete integrability. Third, we give a rigorous version of a formal asymptotic calculation of Rowlands to establish the instability of a class of real-valued periodic waves in 1D, which includes the cnoidal waves of the 1D cubic focusing nonlinear Schrödinger equation, against perturbations with period a large multiple of their fundamental period. Finally, we develop a numerical method to compute the minimizers of the energy with fixed mass and momentum constraints. Numerical experiments support and complete our analytical results.

1 Introduction

We consider the cubic nonlinear Schrödinger equation
(1.1)
in one space dimension, where ψ:R×RC and bR{0}. Equation (1.1) has well-known applications in optics, quantum mechanics, and water waves, and serves as a model for nonlinear dispersive wave phenomena more generally [11, 31]. It is said to be focusing if b>0 and defocusing if b<0. Note that (1.1) is invariant under
  • spatial translation: ψ(t,x)ψ(t,x+a) for aR

  • phase multiplication: ψ(t,x)eiαψ(t,x) for αR.

We are particularly interested in the spatially periodic setting
The Cauchy problem (1.1) is globally well posed in Hloc1PT [7]. We refer to [6] for a detailed analysis of nonlinear Schrödinger equations with periodic boundary conditions. Solutions to (1.1) conserve mass M, energy E, and momentum :
By virtue of its complete integrability, (1.1) enjoys infinitely many higher (in terms of the number of derivatives involved) conservation laws [27], but we do not use them here, in order to remain in the energy space Hloc1, and with the aim of avoiding techniques which rely on integrability.
The simplest non-trivial solutions of (1.1) are the standing waves, which have the form
and so the profile function u (x) must satisfy the ordinary differential equation
(1.2)
We are interested here in those standing waves eiatu(x) whose profiles u (x) are spatially periodic—which we refer to as periodic waves. One can refer to the book [3] for an overview of the role and properties of periodic waves in nonlinear dispersive PDEs.

Non-constant, real-valued, periodic solutions of (1.2) are well known to be given by the Jacobi elliptic functions: dnoidal (dn), cnoidal (cn) (for b>0), and snoidal (sn) (for b<0)—see Section 2 for details. To make the link with Schrödinger equations set on the whole real line, one can see a periodic wave as a special case of infinite train solitons [25, 26]. Another context in which periodic waves appear is when considering the nonlinear Schrödinger equation on a Dumbbell graph [28]. Our interest here is in the stability of these periodic waves against periodic perturbations whose period is a multiple of that of the periodic wave.

Some recent progress has been made on this stability question. By Grillakis–Shatah–Strauss [18, 19] type methods, orbital stability against energy (Hloc1)-norm perturbations of the same period is known for dnoidal waves [2], and for snoidal waves [13] under the additional constraint that perturbations are anti-symmetric with respect to the half-period. In [13], cnoidal waves are shown to be orbitally stable with respect to half-anti-periodic perturbations, provided some condition is satisfied. This condition is verified analytically for small amplitude cnoidal waves and numerically for larger amplitude. Remark here that the results in [13] are obtained in a broader setting, as they are also considering non-trivially complex-valued periodic waves. Integrable systems’ methods introduced in [5] and developed in [15]—in particular conservation of a higher-order functional—are used to obtain the orbital stability of the snoidal waves against Hloc2 perturbations of period any multiple of that of sn.

Our goal in this paper is to further investigate the properties of periodic waves. We follow three lines of exploration. First, we give global variational characterization of the waves in the class of periodic or half-anti-periodic functions. As a corollary, we obtain orbital stability results for periodic waves. Second, we prove the spectral stability of cnoidal, dnoidal, and snoidal waves within the class of functions whose period is the fundamental period of the wave. Third, we prove that cnoidal waves are linearly unstable if perturbations are periodic for a sufficiently large multiple of the fundamental period of the cnoidal wave.

Our first main results concern global variational characterizations of the elliptic function periodic waves as constrained-mass energy minimizers among (certain subspaces of) periodic functions, stated as a series of propositions in Section 3. In particular, the following characterization of the cnoidal functions seems new. Roughly stated (see Proposition 3.4 for a precise statement): 

Theorem 1.1

Let b>0. The unique (up to spatial translation and phase multiplication) global minimizer of the energy, with fixed mass, among half-anti-periodic functions is a (appropriately rescaled) cnoidal function.□

Due to the periodic setting, existence of a minimizer for the problems that we are considering is easily obtained. The difficulty lies within the identification of this minimizer: is it a plane wave, a (rescaled) Jacobi elliptic function, or something else? To answer this question, we first need to be able to decide whether the minimizer can be considered real-valued after a phase change. This is far from obvious in the half-anti-periodic setting of Theorem 1.1, where we use a Fourier coefficients rearrangement argument (Lemma 3.5) to obtain this information. To identify the minimizers, we use a combination of spectral and Sturm–Liouville arguments.

As a corollary of our global variational characterizations, we obtain orbital stability results for the periodic waves. In particular, Theorem 1.1 implies the orbital stability of all cnoidal waves in the space of half-anti-periodic functions. Such orbital stability results for periodic waves were already obtained in [2, 13] as consequences of local constrained minimization properties. Our global variational characterizations provide alternate proofs of these results—see Corollaries 3.9 and 4.7. The orbital stability of cnoidal waves was proved only for small amplitude in [13], and so we extend this result to all amplitude. Remark, however, once more that we are in this paper considering only real-valued periodic wave profiles, as opposed to [13] in which truly complex valued periodic waves were investigated.

Our second main result proves the linear (more precisely, spectral) stability of the snoidal and cnoidal (with some restriction on the parameter range in the latter case) waves against same-period perturbations, but without the restriction of half-period antisymmetry: 

Theorem 1.2

Snoidal waves and cnoidal waves (for a range of parameters) with fundamental period T are spectrally stable against T-periodic perturbations.□

See Theorem 4.1 for a more precise statement. For sn, this is already a consequence of [5, 15], whereas for cn the result was obtained in [21]. The works [5, 15] and [21] both exploit the integrable structure, so our result could be considered an alternate proof which does not uses integrability, but instead relies mainly on an invariant subspace decomposition and an elementary Krein-signature-type argument. See also the recent work [16] for related arguments.

The proof of Theorem 1.2 goes as follows. The linearized operator around a periodic wave can be written as JL, where J is a skew symmetric matrix and L is the self-adjoint linearization of the action of the wave (see Section 4 for details). The operator L is made of two Lamé operators and we are able to calculate the bottom of the spectrum for these operators. To obtain Theorem 1.2, we decompose the space of periodic functions into invariant subspaces: half-periodic and half-anti-periodic, even and odd. Then we analyse the linearized spectrum in each of these subspaces. In the subspace of half-anti-periodic functions, we obtain spectral stability as a consequence of the analysis of the spectrum of L (alternately, as a consequence of the variational characterizations of Section 3). For the subspace of half-periodic functions, a more involved argument is required. We give in Lemma 4.12 an abstract argument relating coercivity of the linearized action L with the number of eigenvalues with negative Krein signature of JL (this is in fact a simplified version of a more general argument [20]). Since we are able to find an eigenvalue with negative Krein signature for JL, spectral stability for half-periodic functions follows from this abstract argument.

Our third main result makes rigorous a formal asymptotic calculation of Rowlands [30] which establishes: 

Theorem 1.3

Cnoidal waves are unstable against perturbations whose period is a sufficiently large multiple of its own.□

This is stated more precisely in Theorem 5.3, and is a consequence of a more general perturbation result, Proposition 5.4, which implies this instability for any real periodic wave for which a certain quantity has the right sign. In particular, the argument does not rely on any integrability (beyond the ability to calculate the quantity in question in terms of elliptic integrals).

Perturbation argument was also used by [14, 15], but our strategy here is different. Instead of relying on abstract theory to obtain the a priori existence of branches of eigenvalues, we directly construct the branch in which we are interested. This is done by first calculating the exact terms of the formal expansion for the eigenvalue and eigenvector at the two first orders, and then obtaining the rigorous existence for the rest of the expansion using a contraction mapping argument. Note that the branch that we are constructing was described in terms of Evans function in [21].

Finally, we complete our analytical results with some numerical observations. Our motivation is to complete the variational characterizations of periodic waves, which was only partial for snoidal waves. We observe: 

Observation 1.4

Let b<0. For a given period, the unique (up to phase shift and translation) global minimizer of the energy with fixed mass and 0 momentum among half-anti-periodic functions is a (appropriately rescaled) snoidal function.□

We have developed a numerical method to obtain the profile ϕ as a minimizer on two constraints, fixed mass and fixed (zero) momentum. We use a heat flow algorithm, where at each time step, the solution is renormalized to satisfy the constraints. Mass renormalization is simply obtained by scaling. Momentum renormalization is much trickier. We define an auxiliary evolution problem for the momentum that we solve explicitly, and plug back the solution we obtain to get the desired renormalized solutions. We first have tested our algorithm in the cases where our theoretical results hold and we have a good agreement between the theoretical results and the numerical experiments. Then, we have performed experiments on snoidal waves which led to Observation 1.4. 

Remark 1.5

As already mentioned, our goal in this work was to avoid using the integrable structure of the equation. However, our results are still limited to the cubic one dimensional case. Indeed, we are using at several steps of our analysis explicit calculations related to the properties of Jacobi elliptic functions. Similar calculations could be performed in the case of the cubic 1D Klein–Gordon equation, which also admits standing waves with cnoidal, dnoidal, and snoidal profiles, but is not completely integrable. As direct arguments, most calculations could probably be replaced by Sturm–Liouville arguments for more general nonlinearities, but there are some key points (like in the concluding proof of Theorem 4.1) that are probably specific to the cubic nonlinearity. Moreover, we expect all of our conclusions to be robust against small perturbations of the nonlinearity.□

The rest of this paper is divided as follows. In Section 2, we present the spaces of periodic functions and briefly recall the main definitions and properties of Jacobi elliptic functions and integrals. In Section 3, we characterize the Jacobi elliptic functions as global constraint minimizers and give the corresponding orbital stability results. Section 4 is devoted to the proof of spectral stability for cnoidal and snoidal waves, whereas in Section 5 we prove the linear instability of cnoidal waves. Finally, we present our numerical method in Section 6 and the numerical experiments in Section 7.

2 Preliminaries

This section is devoted to reviewing the classification of real-valued periodic waves in terms of Jacobi elliptic functions.

2.1 Spaces of periodic functions

Let T>0 be a period. Denote by τT the translation operator
acting on Lloc2(R), and its eigenspaces
for μC with μ=1. Taking μ=1 yields the space of T-periodic functions
while for μ=1 we get the T-anti-periodic functions
For 2kN, letting μ run through the kth roots of unity: ωk=1, and ωj1 for 1j<k, we have
where the decomposition of fPkT is given by
Only the case k=2 is needed here:
(2.1)
Since the reflection R:f(x)f(x) commutes with τT on P2T, we may further decompose into odd and even components in the usual way
to obtain
and so
(2.2)
Each of these subspaces is invariant under (1.1), since
When dealing with functions in PT, we will denote norms such as Lq(0,T) by
and the complex L2 inner product by
(2.3)

2.2 Jacobi elliptic functions

Here we recall the definitions and main properties of the Jacobi elliptic functions. The reader might refer to treatises on elliptic functions (e.g., [24]) or to the classical handbooks [1, 17] for more details.

Given k(0,1), the incomplete elliptic integral of the first kind in the trigonometric form is
and the Jacobi elliptic functions are defined through the inverse of F(·,k):
The relations
(2.4)
follow. For extreme value k=0, we recover trigonometric functions,
while for extreme value k=1, we recover hyperbolic functions:
The periods of the elliptic functions can be expressed in terms of the complete elliptic integral of the first kind
The functions sn and cn are 4K-periodic, while dn is 2K-periodic. More precisely,
The derivatives (with respect to x) of elliptic functions can themselves be expressed in terms of elliptic functions. For fixed k(0,1), we have
(2.5)
from which one can easily verify that sn, cn, and dn are solutions of
(2.6)
with coefficients a,bR for k(0,1) given by
(2.7)
(2.8)
(2.9)

2.3 Elliptic integrals

For k(0,1), the incomplete elliptic integral of the second kind in trigonometric form is defined by
The complete elliptic integral of the second kind is defined as
We have the relations (using dθ=dn(z,k)dz and x=F(ϕ,k))
(2.10)
relating the elliptic functions to the elliptic integral of the second kind, and
(2.11)
relating the elliptic integrals of first and second kind. We can differentiate E and K with respect to k and express the derivatives in terms of E and K:
Note in particular K is increasing, E is decreasing. Moreover,

2.4 Classification of real periodic waves

Here we make precise the fact that the elliptic functions provide the only (non-constant) real-valued, periodic solutions of (2.6). Note that there is a two-parameter family of complex-valued, bounded, solutions for every a,bR, b0 [12, 14]. 

Lemma 2.1 (Focusing case)

Fix a period T>0, aR, b>0 and uPT a non-constant real solution of (2.6). By invariance under translation, and negation (uu), we may suppose u(0)=maxu>0:

  • if 0minu<u(0), then a<0, a<bu(0)2<2a, and u(x)=1αdn(xβ,k),

  • if minu<0, then max(0,2a)<bu(0)2, and u(x)=1αcn(xβ,k),

for some α>0, β>0, and 0<k<1, uniquely determined by T, a, b, and maxu. They satisfy the a-independent relations bβ2=2α2 for (a) and bβ2=2k2α2 for (b). In addition, there exists nN such that 2K(k)βn=T for (a) and 4K(k)βn=T for (b).□

Note that here T may be any multiple of the fundamental period of u. An a-independent relation is useful since a will be the unknown Lagrange multiplier for our constrained minimization problems in Section 3. 

Proof
The first integral associated with (2.6) (written as an Hamiltonian system in x) is constant: there exists C0R such that
A periodic solution has to oscillate in the energy well W(u)=au2+b2u4 with energy level C0. If 0minu, then a<0 and C0<0. If minu<0, then C0>0. Let u(x)=1αv(xβ) with α=(maxu)1. Then v satisfies

(a) If 0minu, then a<0 and C0<0. Let 0<y1<y2 be the roots of ay+b2y2=C0<0. Then u(0)2=y2(a/b,2a/b).

Let β=α2/b. Then bβ2α2=2 and aβ2(2,1), and there is a unique k(0,1) so that aβ2=2+k2. Thus
By uniqueness of the ODE, v(x)=dn(x,k) is the only solution. Hence u(x)=1αdn(xβ,k).
(b) If minu<0, then C0>0. Let y1<0<y2 be the roots of ay+b2y2=C0>0. Then u(0)2=y2>max(0,2a/b) no matter a<0 or a0. We claim we can choose unique β>0 and k(0,1) so that
The sum gives (a+bα2)β2=1, thus β=(a+bα2)1/2 noting (a+bα2)>0, and
no matter a<0 or a0. Thus
By uniqueness of the ODE, v(x)=cn(x,k) is the only solution. Hence, u(x)=1αcn(xβ,k). ■

 

Lemma 2.2 (Defocusing case)

Fix a period T>0, aR, b<0 and uPT a non-constant, real solution of (2.6). By invariance under translation and negation, suppose u(0)=maxu>0. Then 0<bu(0)2<a, and u(x)=1αsn(K(k)+xβ,k), for some α>0, β>0, and 0<k<1, uniquely determined by T, a, b, and maxu. They satisfy the a-independent relation bβ2=2k2α2. In addition, there exists nN such that 4K(k)βn=T.□

 
Proof
The first integral is constant: there exists C0R such that
A periodic solution has to oscillate in the energy well W(u)=au2+b2u4 with energy level C0. Hence, a>0 and 0<C0<maxW=a22b. Let u(x)=1αv(xβ) with α=(maxu)1. Then v satisfies

Let 0<y1<y2 be the roots of ay+b2y2=C0. Then u(0)2=y1(0,a/b).

Let β=(2α22α2a+b)1/2 and k=(b2α2a+b)1/2, noting 2α2a+b>0. Then aβ2=1+k2, bβ2α2=2k2, and v satisfies
By uniqueness of the ODE, v(x)=sn(K(k)+x,k) is the only solution. Hence u(x)=1αsn(K(k)+xβ,k). ■

3 Variational Characterizations and Orbital Stability

Our goal in this section is to characterize the Jacobi elliptic functions as global constrained energy minimizers. As a corollary, we recover some known results on orbital stability, which is closely related to local variational information.

3.1 The minimization problems

Recall the basic conserved functionals for (1.1) on Hloc1PT:
In this section, we consider L2(0,T;C) as a real Hilbert space with scalar product Re0Tfg¯dx. The space H1 is also considered as a real Hilbert space. This way, the functionals E, M, and are C1 functionals on H1. This also ensures that the Lagrange multipliers are real. Note that we see L2(0,T;C) as a real Hilbert space only in the current section and in all the other sections, it will be seen as a complex Hilbert space with the scalar product defined in (2.3).
Fix parameters T>0, a,bR, b0. Since the Jacobi elliptic functions (indeed any standing wave profiles) are solutions of (2.6), they are critical points of the action functional Sa defined by
where the values of a and b are given in (2.7)(2.9) and the fundamental periods are T=2K for dn, T=4K for sn,cn. Given m>0, the basic variational problem is to minimize the energy with fixed mass:
(3.1)
whose Euler–Lagrange equation
(3.2)
with aR arising as Lagrange multiplier, is indeed of the form (2.6). Since the momentum is also conserved for (1.1), it is natural to consider the problem with a further momentum constraint:
(3.3)
 
Remark 3.1

Note that if a minimizer u of (3.1) is such that P(u)=0, then it is real-valued (up to multiplication by a complex number of modulus 1). Indeed, it verifies (3.2) for some aR. It is well known (see e.g., [13]) that the momentum density I(uxu¯) is, therefore, constant in x, and so it is identically 0 if (u)=0. For u(x)0, we can write u as u=ρeiθ, and express the momentum density as Im(uxu¯)=θxρ2. Thus Im(uxu¯)=0 implies θx=0 and thus θ(x) is a constant as long as u(x)0. If u(x0)=0 and eθ(x0)eθ(x0+), we must have ux(x0)=0, and hence u0 by uniqueness of the ODE.□

Since (1.1) preserves the subspaces in the decomposition (2.1), it is also natural to consider variational problems restricted to anti-symmetric functions
(3.4)
(3.5)
and in light of the decomposition (2.2), further restrictions to even or odd functions may also be considered.

In general, the difficulty does not lie in proving the existence of a minimizer, but rather in identifying this minimizer with an elliptic function, since we are minimizing among complex valued functions, and moreover restrictions to symmetry subspaces prevent us from using classical variational methods like symmetric rearrangements.

We will first consider the minimization problems (3.1) and (3.3) for periodic functions in PT. Then we will consider the minimization problems (3.4) and (3.5) for half-anti-periodic functions in AT/2. In both parts, we will treat separately the focusing (b>0) and defocusing (b<0) nonlinearities. For each case, we will show the existence of a unique (up to phase shift and translation) minimizer, and we will identify it with either a plane wave or a Jacobi elliptic function.

3.2 Minimization among periodic functions

3.2.1 The focusing case in PT

 

Proposition 3.2

Assumeb>0. The minimization problems(3.1)and(3.3)satisfy the following properties:

  • For allm>0, (3.1)and(3.3)share the same minimizers. The minimal energy is finite and negative.

  • For all0<mπ2bT, there exists a unique (up to phase shift) minimizer of(3.1). It is the constant functionumin2mT.

  • For allπ2bT<m<, there exists a unique (up to translations and phase shift) minimizer of(3.1). It is the rescaled functiondnα,β,k=1αdn·β,k, where the parameters α, β, and k are uniquely determined. Its fundamental period is T. The map fromm(π2bT,)tok(0,1)is one-to-one, on to, and increasing.

  • In particular, givenk(0,1), dn=dn(·,k), ifb=2, T=2K(k), andm=M(dn)=E(k), then the unique (up to translations and phase shift) minimizer of(3.1)isdn.□

 
Proof
Without loss of generality, we can restrict the minimization to real-valued non-negative functions. Indeed, if uHloc1PT, then uHloc1PT and we have
This readily implies that (3.1) and (3.3) share the same minimizers. Let us prove that
(3.6)
The last inequality in (3.6) is obtained using the constant function φm,02mT as a test function:
To prove the first inequality in (3.6), we observe that by Gagliardo–Nirenberg inequality, we have
Consequently, for uHloc1PT such that M(u)=m, we have
and E has to be bounded from below. The above shows (i).
Consider now a minimizing sequence (un)Hloc1PT for (3.1). It is bounded in Hloc1PT and, therefore, up to a subsequence, it converges weakly in Hloc1PT and strongly in Lloc2PT and Lloc4PT towards uHloc1PT. Therefore, E(u)E(un) and M(u)=m. This implies that xuL2=limnxunL2 and, therefore, the convergence from un to u is also strong in Hloc1PT. Since u is a minimizer of (3.1), there exists a Lagrange multiplier aR such that
that is
Multiplying by u and integrating, we find that
Note that
therefore,
We already have uR, and we may assume maxu=u(0) by translation. By Lemma 2.1 (a), either u is constant or there exist α,β(0,) and k(0,1) such that β=α2/b and
We now show that the minimizer u is of the form dnα,β,k if m>π2bT. Indeed, assuming by contradiction that u is a constant, we necessarily have u2mT. The Lagrange multiplier can also be computed and we find a=bu2=2bmT. Since u is supposed to be a constrained minimizer for (3.1), the operator
must have Morse index at most 1, i.e., at most 1 negative eigenvalue. The eigenvalues are given for nZ by the following formula:
Obviously, n=0 gives a negative eigenvalue. For n=1, the eigenvalue is non-negative if and only if
which gives the contradiction. Hence, when m>π2bT, the minimizer u must be of the form dnα,β,k.
There is a positive integer n so that the fundamental period of u=dnα,β,k is 2K(k)β=Tn1. As already mentioned, since u is a minimizer for (3.1), the operator
can have at most one negative eigenvalue. The function xu is in its kernel and has 2n zeros. By Sturm–Liouville theory (see e.g., [10, 29]), we have at least 2n1 eigenvalues below 0. Hence, n=1 and 2K(k)β=T.
Using 2α2=bβ2 (see Lemma 2.1), the mass verifies,
where E (k) is given in Section 2.3. Using 2K(k)β=T,
(3.7)
Note
where the positivity of the numerator is because it vanishes at k=0 and
Thus EK (k) varies from π24 to when k varies from 0 to 1. Thus (3.7) defines m as a strictly increasing function of k(0,1) with range (π2bT,) and hence has an inverse function. For fixed b,m,T, the value k(0,1) is uniquely determined by (3.7). We also have β=T2K(k) and α=βb/2. The above shows (iii).

The above calculation also shows that m>π2bT if u=dnα,β,k. Thus, u must be a constant when 0<mπ2bT. This shows (ii).

In the case we are given k(0,1), T=2K(k), b=2 and m=M(dn)=E(k), we want to show that u(x)=dn(x,k). In this case, m>π2bT since EK>π24. Thus, by Lemma 2.1 (a), u=dnα,β,s for some α,β>0 and s(0,1), up to translation and phase. By the same Sturm–Liouville theory argument, the fundamental period of u is T=2K(s)β. The same calculation leading to (3.7) shows
Thus, E(k)K(k)=E(s)K(s). Using the monotonicity of EK (k) in k, we have k=s. Thus α=β=1 and u(x)=dn(x,k). This gives (iv) and finishes the proof.■

3.2.2 The defocusing case in PT

 

Proposition 3.3

Assumeb<0. For all0<m<, the constrained minimization problems(3.1)and(3.3)have the same unique (up to phase shift) minimizers, which is the constant functionumin2mT.□

 
Proof
This is a simple consequence of the fact that functions with constant modulus are the optimizers of the injection L4(0,T)L2(0,T). More precisely, for every fL4(0,T), we have by Hölder's inequality,
with equality if and only if f is constant. Let φm,0 be the constant function φm,02mT. For any vHloc1PT such that M(v)=m and veiθφm,0 (θR), we have
As a consequence, E(φm,0)<E(v) and this proves the proposition.■

3.3 Minimization among half-anti-periodic functions

3.3.1 The focusing case in AT/2

Proposition 3.4
 

Assumeb>0. For allm>0, the minimization problems(3.4)and(3.5)inAT/2satisfy the following properties:

  • The minimizers for(3.4)and(3.5)are the same.

  • There exists a unique (up to translations and phase shift) minimizer of(3.4). It is the rescaled functioncnα,β,k=1αcn·β,k, where the parameters α, β, and k are uniquely determined. Its fundamental period is T. The map fromm(0,)tok(0,1)is one-to-one, on to, and increasing.

  • In particular, givenk(0,1), cn=cn(·,k), ifb=2k2, T=4K(k), andm=M(cn)=2(E(1k2)K)/k2, then the unique (up to translations and phase shift) minimizer of(3.4)iscn.□

Before proving Proposition 3.4, we make the following crucial observation. 

Lemma 3.5
LetvHloc1AT/2. Then there existsv˜Hloc1AT/2such that
 
Proof of Lemma 3.5
The proof relies on a combinatorial argument. Since vHloc1AT/2, its Fourier series expansion contains only terms indexed by odd integers:
We define v˜ by its Fourier series expansion
It is clear that v˜(x)R for all xR, and by Plancherel formula,
so all we have to prove is that v˜L4vL4. We have
where we have defined
Using the fact that for nN, n0, the term ein2πTx integrates to 0 due to periodicity,
we compute
The first part is just
For the second part, we observe that
(3.8)
where the · denotes the complex vector scalar product. Therefore,
where by w˜n we denote the quantity defined similarly as in (3.8) for (v˜j). As a consequence,
and this finishes the proof of Lemma 3.5.■
 
Proof of Proposition 3.4

All functions are considered in AT/2. Consider a minimizing sequence (un) for (3.5). By Lemma 3.5, the minimizing sequence can be chosen such that un(x)R for all xR and this readily implies the equivalence between (3.5) and (3.4), which is (i).

Using the same arguments as in the proof of Proposition 3.2, we infer that the minimizing sequence converges strongly in Hloc1AT/2 to uHloc1AT/2 verifying for some aR the Euler–Lagrange equation
Then, since u is real and in AT/2, we may assume maxu=u(0)>0 and, by Lemma 2.1 (b), there exists a set of parameters α,β(0,), k(0,1) such that
and the parameters α,β,k are determined by T, a, b, and maxu, with 2k2α2=bβ2.
There exists an odd, positive integer n so that the fundamental period of u is 4K(k)β=T/n (n has to be odd, otherwise, if n=2k with kN, u would be periodic of period kT/n=T/2, which is not possible since u is in AT/2). Since u is a minimizer for (3.4), the operator
can have at most one negative eigenvalue in Lloc2AT/2. The function xu is in its kernel and has 2n zeros in [0,T). By Sturm–Liouville theory, there are at least n1 eigenvalues (with eigenfunctions in AT/2) below 0. Hence, since n is odd, n = 1 and 4K(k)β=T.
The mass verifies, using 2k2α2=bβ2 and (2.11),
Using 4K(k)β=T,
(3.9)
Note all factors of M (k) are positive, kK(k)>0 and
Thus, (3.9) defines m as a strictly increasing function of k(0,1) with range (0,) and hence has an inverse function. For fixed T,b,m, the value k(0,1) is uniquely determined by (3.9). We also have β=T4K(k) and α2=bβ22k2. The above shows (ii).
In the case, we are given k(0,1), T=4K(k), b=2k2 and m=M(cn(·,k)), we want to show that u(x)=cn(x,k). In this case, by Lemma 2.1 (b), u=cnα,β,s for some α,β>0 and s(0,1), up to translation and phase. By the same Sturm–Liouville theory argument, the fundamental period of u is T=4K(s)β. The same calculation leading to (3.9) shows
Thus, M(s)=M(k). By the monotonicity of M (k) in k, we have k = s. Thus, α=β=1 and u(x)=cn(x,k). This shows (iii) and concludes the proof.■

3.3.2 The defocusing case in AT/2

 

Proposition 3.6

Assume b<0. There exists a unique (up to phase shift and complex conjugate) minimizer for (3.4). It is the plane wave umin2mTe2iπxT.□

 
Proof
Denote the supposed minimizer by w(x)=2mTe±2iπxT. Let vHloc1A2K such that M(v)=m and veiθw (θR). As in the proof of Proposition 3.3, we have
Since vA2K, v must have 0 mean value. Recall that in that case v verifies the Poincaré–Wirtinger inequality
and that the optimizers of the Poincaré–Wirtinger inequality are of the form Ce±2iπTx, CC. This implies that
As a consequence, E(w)<E(v) and this proves the lemma.■

As far as (3.5) is concerned, we make the following conjecture 

Conjecture 3.7

Assume b<0. The unique (up to translations and phase shift) minimizer of (3.5) is the rescaled function snα,β,k=1αsn·β,k, where the parameters α, β, and k are uniquely determined.

In particular, given k(0,1), sn=sn(·,k), ifb=2k2, T=4K(k), and m=M(sn), then the unique (up translations and to phase shift) minimizer of (3.5) is sn.□

This conjecture is supported by numerical evidence, see Observation 7.1. The main difficulty in proving the conjecture is to show that the minimizer is real up to a phase.

3.3.3 The defocusing case in AT/2

In light of our uncertainty about whether sn solves (3.5), let us settle for the simple observation that it is the energy minimizer among odd, half-anti-periodic functions: 

Proposition 3.8
Assume b<0. The unique (up to phase shift) minimizer of the problem
(3.10)
is the rescaled function snα,β,k=1αsn·β,k, where the parameters α, β, and k are uniquely determined. Its fundamental period is T. The map from m(0,) to k(0,1) is one-to-one, on to, and increasing.

In particular, given k(0,1), sn=sn(·,k), if b=2k2, T=4K(k), and m=M(sn), then the unique (up to phase shift) minimizer of (3.10) is sn.□

 
Proof
If uAT/2, then 0=u(0)=u(T/2), and since u is completely determined by its values on [0,T/2], we may replace (3.10) by
for which the map uu is admissible, showing that minimizers are non-negative (up to phase), and in particular real-valued, hence, a (rescaled) sn function by Lemma 2.2. The remaining statements follow as in the proof of Proposition 3.4. In particular, the mass verifies, using 2k2α2=bβ2, (2.11), and 4K(k)β=T,
which is a strictly increasing function of k(0,1) with range (0,) and hence has an inverse function.■

3.4 Orbital stability

Recall that we say that a standing wave ψ(t,x)=eiatu(x) is orbitally stable for the flow of (1.1) in the function space X if for all ε>0, there exists δ>0 such that the following holds: if ψ0X verifies
then the solution ψ of (1.1) with initial data ψ(0,x)=ψ0 verifies for all tR the estimate
As an immediate corollary of the variational characterizations above, we have the following orbital stability statements: 
Corollary 3.9
The standing wave ψ(t,x)=eiatu(x) is a solution of (1.1), and is orbitally stable in X in the following cases. For Jacobi elliptic functions: for any k(0,1),
For constants and plane waves: (b0)

The proof of this corollary uses the variational characterizations from Propositions 3.2, 3.3, 3.4, 3.6, and 3.8. Note that for all the minimization problems considered, we have the compactness of minimizing sequences. The proof follows the standard line introduced by Cazenave and Lions [8], we omit the details here. 

Remark 3.10

The orbital stability of sn [13] in Hloc1AT/2 was proved using the Grillakis–Shatah–Strauss [18, 19] approach, which amounts to identifying the periodic wave as a local constrained minimizer in this subspace. So the above may be considered an alternate proof, using global variational information. In the case of sn, without Conjecture 3.7, some additional spectral information in the subspace AT/2+ is needed to obtain orbital stability in Hloc1AT/2 (rather than just Hloc1AT/2)—see Corollary 4.7 in the next section for this.

Orbital stability of cn was obtained in [13] only for small amplitude cn. We extend this result to all possible values of k(0,1).□

 
Remark 3.11

Using the complete integrability of (1.1), Bottman et al. [5], and Gallay and Pelinovsky [15] showed that sn is in fact a minimizer of a higher-order functional in Hloc2PnT for any n, and thus showed it is orbitally stable in these spaces.□

4 Spectral stability

Given a standing wave ψ(t,x)=eiatu(x) solution of (1.1), we consider the linearization of (1.1) around this solution: if ψ(t,x)=eiat(u(x)+h), then h verifies
where L denotes the linear part and N the nonlinear part. Assuming u is real-valued, we separate h into real and imaginary parts to get the equation
where
We call
(4.1)
the linearized operator of (1.1) about the standing wave eiatu(x).

Now suppose uHloc1PT is a (period T) periodic wave, and consider its linearized operator JL as an operator on the Hilbert space (PT)2, with domain (Hloc2PT)2. The main structural properties of JL are:

  • since L± are self-adjoint operators on PT, L is self-adjoint on (PT)2, while J is skew-adjoint and unitary
    (4.2)
  • JL commutes with complex conjugation,
    (4.3)
  • JL is antisymmetric under conjugation by the matrix
    (which corresponds to the operation of complex conjugation before complexification),
    (4.4)

At the linear level, the stability of the periodic wave is determined by the location of the spectrum σ(JL), which in this periodic setting consists of isolated eigenvalues of finite multiplicity [29]. We first make the standard observation that as a result of (4.3) and (4.4), the spectrum of JL is invariant under reflection about the real and imaginary axes:
Indeed, if JLf=λf, then
We are interested in whether the entire spectrum of JL lies on the imaginary axis, denoted σ(JLPT)iR, in which case, we say the periodic wave u is spectrally stable in PT. Moreover, if SPT is an invariant subspace—more precisely, JL:(Hloc2S)2(S)2—then we will say that the periodic wave u is spectrally stable in S if the entire (S)2 spectrum of JL lies on the imaginary axis, denoted σ(JLS)iR. In particular, for k(0,1) and K=K(k), since sn2, cn2, dn2P2K+, the corresponding linearized operators respect the decomposition (2.2), and we may consider σ(JLS) for S=P2K±,A2K±P4K, with
(4.5)

Of course, spectral stability (which is purely linear) is a weaker notion than orbital stability (which is nonlinear). Indeed, the latter implies the former—see Proposition 4.10 and the remarks preceding it.

The main result of this section is the following. 

Theorem 4.1

Spectral stability in PT, T=4K(k), holds for:

  • u=sn, k(0,1),

  • u=cn and k(0,kc), where kc is the unique k(0,1) so that K(k)=2E(k), kc0.908.□

 
Remark 4.2

The function f(k)=K(k)2E(k) is strictly increasing in k(0,1) (since K (k) is increasing while E (k) is decreasing in k), with f(0)=π2 and f(1)=.□

 
Remark 4.3

Using Evans function techniques, it was proved in [21] that σ(JLcn)iR also for k[kc,1). This fact is also supported by numerical evidence (see Section 7).□

 
Remark 4.4

In the case of sn, the Hloc2PnT orbital stability obtained in [5, 15] (using integrability) immediately implies spectral stability in PnT, and in particular in PT. So our result for sn could be considered an alternate, elementary proof, not relying on the integrability.□

 
Remark 4.5

The spectral stability of dn in P2K (its own fundamental period) is an immediate consequence of its orbital stability in Hloc1P2K, see Proposition 4.10.□

4.1 Spectra of L+ and L

We assume now that we are given k(0,1) and we describe the spectrum of L+ and L in P4K when ϕ is cn, dn or sn. When ϕ=sn, we denote L+ by L+sn, and we use similar notations for L and cn,dn. Due to the algebraic relationships between cn, dn, and sn, we have
Similarly for L, we obtain
As a consequence, L±sn, L±cn, and L±dn share the same eigenvectors. Moreover, these operators enter in the framework of Schrödinger operators with periodic potentials and much can be said about their spectrum (see e.g., [10, 29]). Recall in particular that given a Schrödinger operator L=xx+V with periodic potential V of period T, the eigenvalues λn of L on PT satisfy
with corresponding eigenfunctions ψn such that ψ0 has no zeros, ψ2m+1 and ψ2m+2 have exactly 2m+2 zeros in [0,T) ([10], p. 39). From the equations satisfied by cn, dn, sn, we directly infer that
Taking the derivative with respect to x of the equations satisfied by cn, dn, sn, we obtain
Looking for eigenfunctions in the form χ=1Asn2 for AR, we find two other eigenfunctions:
where
In the interval [0,4K), χ has no zero, snx and cnx have two zeros each, while dnx and χ+ have 4 zeros each. By Sturm–Liouville theory, they are the first five eigenvectors of L+ for each of sn, cn, and dn, and all other eigenfunctions have strictly greater eigenvalues Similarly, dn>0 has no zeros, while cn and sn have two each, so these are the first three eigenfunctions of L for each of sn, cn, and dn, and all other eigenfunctions have strictly greater eigenvalues.

The spectra of L±sn, L±cn, and L±dn are represented in Figure 1, where the eigenfunctions are also classified with respect to the subspaces of decomposition (2.2).

Eigenvalues for L− and L+ in P4K.
Fig. 1.

Eigenvalues for L and L+ in P4K.

We may now recover the result of [13] that sn is orbitally stable in Hloc1A2K, using the following simple consequences of the spectral information above: 

Lemma 4.6

There exists δ>0 such that the following coercivity properties hold:

  1. L+snA2K>δ,

  2. LsnA2K{sn}>δ,

  3. L+snA2K+{(sn)x}>δ,

  4. LsnA2K+{(sn)x}>δ.□

 
Proof
The first three are immediate from Figure 1 (note the first two also follow from the minimization property Proposition 3.8), while we see that in A2K+, L+sn{(sn)x}>e+, so since sn2(x)1,
where the last inequality is easily verified.■
 
Corollary 4.7

For all k(0,1), the standing wave ψ(t,x)=ei(1+k2)tsn(x,k) is orbitally stable in Hloc1A2K.□

 
Proof

Lemma 4.6 shows that sn is a non-degenerate (up to phase and translation) local minimizer of the energy with fixed mass and momentum. So the classical Cazenave–Lions [8] argument yields the orbital stability.■

Finally, we also record here the following computations concerning L±cn, used in analyzing the generalized kernel of JLcn in the next subsection: 

Lemma 4.8
Define Eˆ(x,k)=E(ϕ,k)sinϕ=sn(x,k). Let ϕ1 and ξ1 be given by the following expressions:
The denominators are positive and we have
Note Eˆ and ξ1 are odd, while ϕ1 is even. In particular, (ϕ1,cnx)=0=(ξ1,cn). Moreover, L+cn(12cn(12k2)ϕ1)=cnxx.□
 
Proof
Recall that the elliptic integral of the second kind Eˆ(x,k) is not periodic. In fact, it is asymptotically linear in x and verifies
By (2.10), xEˆ(x,k)=dn2(x,k). Denote L±=L±cn in this proof. Using (2.4) and (2.5), we have
Define
Then ϕ˜1 is periodic (of period 4K) and verifies
The factor is positive if 2k21. If 2k2<1, it is greater than 2(2k21)+2(1k2)=2k2. Define,
Then
As for L, we have
Define
Then ξ˜1 is periodic (of period 4K) and verifies
The factor is positive by (2.11). Defining
we get Lξ1=cnx. The last statement of the lemma follows from (2.8).■

4.2 Orthogonality properties

The following lemma records some standard properties of eigenvalues and eigenfunctions of the linearized operator JL, which follow only from the structural properties (4.2) and (4.4): 

Lemma 4.9

The following properties hold:

  1. (symplectic orthogonality of eigenfunctions) Let f=(f1,f2)T and g=(g1,g2)T be two eigenvectors of JL corresponding to eigenvalues λ,μC. Then (4.2) implies
    while (4.4) implies
    so that
  2. (unstable eigenvalues have zero energy) If JLf=λf, λiR, then (4.2) implies

 
Proof
We first prove (1). We have
so (λ+μ¯)(f,Jg)=0 which gives the first statement. The second statement follows from the same argument with f replaced by Cf, while the third statement is a consequence of (f,Jg)=(Cf,Jg)=0.

Item (2) is a special case of the first statement of (1), with g=f. ■

4.3 Spectral stability of sn and cn

Our goal in this section is to establish Theorem 4.1, i.e., to prove the spectral stability of sn in P4K for all k(0,1), and the spectral stability of cn in P4K for all k(0,kc).

We first recall the standard fact that
Indeed, an eigenvalue λ=α+iβ of JL with α>0 produces a solution of the linearized equation whose magnitude grows at the exponential rate eαt, and this linear growing mode (together with its orthogonality properties from Lemma 4.9) can be used to contradict orbital stability. Rather than go through the nonlinear dynamics, however, we will give a simple direct proof of spectral stability in the symmetry subspaces where we have the orbital stability—that is, in P2K for dn, and in A2K for cn and sn—using just the spectral consequences for L± implied by the (local) minimization properties of these elliptic functions: 
Proposition 4.10
For 0<k<1, K=K(k), dn is spectrally stable in P2K, while cn and sn are spectrally stable in A2K. Precisely, we have
 
Proof
Begin with dn in P2K. From Figure 1, we see Ldndn>0, and thus (Ldn)±1/2 exist on dn. It follows from the minimization property Proposition 3.2 that on dn, L+dn0 (otherwise there is a perturbation of dn lowering the energy while preserving the mass). Suppose JLdnf=λf, λiR. Then LdnL+dnf1=λ2f1. Since (dn,0)T is an eigenvector of JL for the eigenvalue 0, Lemma 4.9 implies f1dn. Therefore, we have
and on dn,
contradicting λiR.

Next, consider cn in A2K. Again from Figure 1, we see Lcncn>0, while the minimization property Proposition 3.4 implies that L+cn0 on cn, and so the spectral stability follows just as for dn above.

Finally, consider sn in A2K. By Lemma 4.6, L+sn>0 on {(sn)x}, while Lsn0 on {(sn)x}, and so the spectral stability follows from the same argument as above, with the roles of L+ and L reversed.■

Moreover, both sn and cn are spectrally stable in P2K: 

Lemma 4.11
For 0<k<1, K=K(k),
 
Proof

This is an immediate consequence of the positivity of Lsn and Lcn on P2K (see Figure 1), and Lemma 4.9.■

So in light of (4.5), to prove Theorem 4.1, it remains only to show σ(JLP2K+)iR for each of cn and sn.

This will follow from a simplified version of a general result for infinite dimensional Hamiltonian systems (see [20, 22, 23]) relating coercivity of the linearized energy with the number of eigenvalues with negative Krein signature of the linearized operator JL of the form (4.1): 

Lemma 4.12 (Coercivity lemma)
Consider JL on S×S for some invariant subspace SPT, and suppose it has an eigenvalue whose eigenfunction ξ=(ξ1,ξ2)T has negative (linearized) energy:
Then the following results hold:
  1. If L+ has a 1D negative subspace (in S):
    (4.6)
    then L+ξ2>0.
  2. If L has a 1D negative subspace (in S):
    (4.7)
    then Lξ1>0.
  3. If both (4.6) and (4.7) hold, then σ(JLS×S)iR.□

 
Proof

First note that by Lemma 4.9 (2), 0μiR, and writing μ=iγ, 0γR, we have Lξ2=iγξ1, L+ξ1=iγξ2.

Moreover,
so by assumption (ξ1,L+ξ1)=(ξ2,Lξ2)<0.
We prove (1). For any hξ2, decompose
where we may assume α0 and β0. We have
Thus, using L+f>0, L+1/2=(L+f)1/2 is well defined on f and
with both factors on the right >0. Since (ξ1,L+ξ1)<0, we must have (h,L+h)>0.

Statement (2) follows in exactly the same way, with the roles of L+ and L reversed, the roles of ξ1 and ξ2 reversed, and with g and ν replacing f and λ.

Finally, for (3), suppose JLη=ζη. If ζiR, then by Lemma 4.9 (1), (ξ1,η2)=(ξ2,η1)=0, and so by parts (1) and (2),
contradicting Lemma 4.9 (2). Thus, ζiR. ■
 
Proof of Theorem 4.1
Begin with sn in P2K+. From Figure 1, it is clear that in P2K+, condition (4.6) holds for L+sn and (4.7) holds for Lsn. Explicit computation yields
which implies
Moreover,
by (2.11). Hence, all the conditions of Lemma 4.12 are verified for sn in P2K+, and so we conclude σ(JLsnP2K+)iR, as required.
Next we turn to cn. Again from Figure 1, it is clear that in P2K+, condition (4.6) holds for L+cn and (4.7) holds for Lcn. Explicit computation yields
which implies
Moreover, when k<kc, we have
Hence, the conditions of Lemma 4.12 are verified for cn in P2K+ when k<kc, yielding σ(JLcnP2K+)iR, as required.■

5 Linear instability

Theorem 4.1 (and Proposition 4.10) give the spectral stability of the periodic waves dn, sn, and cn (at least for k<kc) against perturbations which are periodic with their fundamental period. It is also natural to ask if this stability is maintained against perturbations whose period is a multiple of the fundamental period. In light of Bloch–Floquet theory, this question is also relevant for stability against localized perturbations in L2(R).

5.1 Theoretical analysis

It is a simple observation that dn immediately becomes unstable against perturbations with twice its fundamental period: 

Proposition 5.1

Both σ(JLdnA2K+) and σ(JLdnA2K) contain a pair of non-zero real eigenvalues. In particular dn is linearly unstable against perturbations in P4K.□

 
Proof
In each of A2K+ and A2K, Ldn>0 while L+dn has a negative eigenvalue: L+dnf=λf, λ>0. So the self-adjoint operator (Ldn)1/2L+dn(Ldn)1/2 has a negative direction,
hence, a negative eigenvalue (Ldn)1/2L+dn(Ldn)1/2g=μ2g, μ>0. Setting h(Ldn)1/2g, hA2K+ (A2K), we see
Hence, μ,μR are eigenvalues of JL in A2K+ (A2K).■
 
Remark 5.2

The proof shows dn is unstable in P2nK for every even n since hP2nK. In fact, dn is unstable in any P2nK, n2. Indeed, we always have Ldn=0, thus by Sturm–Liouville Theory (see e.g., [10], Theorem 3.1.2), 0 is always the first simple eigenvalue of L in P2nK. Moreover, L+dnx=0, and dnx has 2n zeros in P2nK. Hence, there are at least 2n2 negative eigenvalues for L+ in P2nK. With the above argument, this proves linear instability in P2nK for any n2.□

For sn, the H2(R) orbital stability result of [5, 15] implies spectral stability against perturbations which are periodic with any multiple of the fundamental period.

Using formal perturbation theory, [30] showed that cn becomes unstable against perturbations which are periodic with period a sufficiently large multiple of the fundamental period. Our main goal in this section is to make this rigorous: 

Theorem 5.3

For 0<k<1, there exists n1=n1(k)N such that cn is linearly unstable in P4nK for nn1, i.e., the spectrum of JLcn as an operator on P4nK contains an eigenvalue with positive real part.□

We will in fact prove a slightly more general result, which is the existence of a branch strictly contained in the first quadrant for the spectrum of JLcn considered as an operator on L2(R). Theorem 5.3 will be a consequence of a more general perturbation result applying to all real periodic waves (see Proposition 5.4), and in particular not relying on any integrable structure.

We start with some preliminaries. Let
with
where u a periodic solution to
(5.1)
We assume that u(x)R and let T denote a period of u2. The spectrum of JL as an operator on L2(R) can be analyzed using Bloch–Floquet decomposition. For θ[0,2π/T), define
where L±θ is the operator obtained when formally replacing x by x+iθ=eiθxx(eiθx·) in the expression of L±. If we let (Mθf)(x)=eiθxf(x), then L±θ=MθL±Mθ. Then we have
(5.2)
Remark here that when θ=πnT for some nN, then we have
(5.3)
In what follows, all operators are considered on PT unless otherwise mentioned.
Let us consider the case θ=πT. Denote
Since u is a real-valued periodic solution to (5.1), by Lemmas 2.1 and 2.2, u is a rescaled cn, dn, or sn. In any case, the following holds:
(5.4)
Note that for any f,gHloc1PT, we can integrate by parts with D:
Remark that
Therefore, there exist φ1,ψ1 such that
The kernel of the operator JLπT is generated by ψ0,0φ. On top of that, the generalized kernel of JLπT contains (at least) 0ψ1,φ10.
Our goal is to analyze the spectrum of the operator JLπTε when ε is small. In particular, we want to locate the eigenvalues generated by perturbation of the generalized kernel of JLπT. For the sake of simplicity in notation, when θ=πT, we use a tilde to replace the exponent πT. In particular, we write
 
Proposition 5.4
Assume the condition (5.17) stated below. There exist λ1C with Re(λ1)>0, Im(λ1)>0; b0C; and ε0>0, such that for all 0ε<ε0, there exist λ2(ε)C, b1(ε)C, v2(ε),w2(ε)Hloc2PT,
(5.5)
verifying the following property. Set
(5.6)
(5.7)
Here, L˜+1 is taken from ψ to ψ. Define
Then

Note that the orthogonality conditions in (5.5) are reasonable: the eigenvector is normalized by Pφw=w0=φ, and hence w2φ. To impose v2ψ, we allow b1(ε)ψ in v1 to be ε-dependent to absorb Pψ(vv0). 

Proof of Proposition 5.4
Let us write the expansion of the operators in ε. We have
Therefore,
We expand in ε the equation JLπTελIvw=0 and show that it can be satisfied at each order of ε.
At order O(1), we have
which is satisfied because v0w0ker(JL˜) by definition.
At order O(ε), we have
which can be rewritten, using the expression of v0, w0, and Dφ=ψ, as
(5.8)
(5.9)
It is clear that the functions v1(ε) and w1 defined in (5.7) and (5.6) satisfy (5.8) and (5.9).
At order O(ε2), we consider the equation as a whole, involving also the higher orders of ε. We have
in other words
(5.10)
(5.11)
where
(5.12)
Note that V2 and W2 depend on b0,λ1 and b1,λ2, whereas V3,V4 and W3,W4 also depend on v2 and w2. Our strategy to solve the system (5.10) and (5.11) is divided into two steps. We first ensure that it can be solved at the main order by ensuring that the compatibility conditions
(5.13)
are satisfied. This is achieved by making a suitable choice of b0,λ1. Then, we solve for b1,λ2,v2,w2 by using a Lyapunov–Schmidt argument.
We rewrite the compatibility conditions (5.13) in the following form, using the expressions for v0, w0, v1, and w1, and the properties of φ and ψ:
These equations do not depend on b1 or λ2 although W2 and V2 do. For a moment, we write these equations as
(5.14)
(5.15)
where
Multiplying (5.14) by C2+A2λ12, (5.15) by Bλ1, and subtracting gives
(5.16)
a quadratic equation in λ12 with real coefficients. If A1A20, the roots of (5.16) are given by
We now assume that the discriminant of this quadratic is negative:
(5.17)
which implies that A1A20, and, moreover, guarantees the existence of a root λ1 of (5.16) strictly contained in the first quadrant: Reλ1>0 and Imλ1>0 (the other roots being λ1, ±λ¯1). It follows from (5.17) that B0, and so we may solve (5.14) and set
so that both (5.14) and (5.15) are satisfied.
We now solve for b1,λ2,v2,w2 using a Lyapunov–Schmidt argument. The first step is to solve, given (b1,λ2), projected versions of (5.10) and (5.11),
(5.18)
to obtain v2=v2(b1,λ2)ψ, w2=w2(b1,λ2)φ:
 
Lemma 5.5
Given any b1C, λ2C with b1+λ2M, there is a unique solution
of (5.18), with v2H2+w2H2C(M).□
 
Proof
By the expressions (5.12), we may rewrite system (5.18) as a linear system of v2 and w2,
where
(5.19)
Note that Sv,Sw,Rv, and Rw do not contain v2,w2 or ε. Recalling the definition
it follows from (5.4) that
is bounded, and hence, so is L˜ε1, uniformly in ε for ε sufficiently small, with
Thus
(5.20)
gives (v2(b1,λ2),w2(b1,λ2)) as desired.■

The second step is to plug (v2(b1,λ2),w2(b1,λ2)) back into V3, V4, W3, W4, and solve, for (b1,λ2), the remaining compatibility conditions
(5.21)
which, together with (5.18), complete the solution of the eigenvalue problem. Using (5.20) and (5.12), we may write (5.21) as the system
and then by the expressions (5.7) and (5.6) and (5.19), we may further rewrite as
(5.22)
where Φ is a rational vector function of b1,λ2, and ε; F is a fixed (independent of (b1,λ2))F1; and M=Φ(b1,λ2)ε=0 is the matrix
where in the last step, we used (5.15). The determinant of M is, using (5.15) and (5.14) to eliminate b0,
Since A1,A2,C1,C2,B2 are real, and λ1A1A20, we have detM0, otherwise λ12R.

Thus, (b1,λ2) may be solved from (5.22) for ε sufficiently small by the implicit function theorem, providing the required solution to (5.21), and so completing the proof of Proposition 5.4.■ 

Proof of Theorem 5.3
We need only verify the assumptions of Proposition 5.4 for the case of u(x)=cn(x;k), T=2K(k). Since u=cnA2K, we have u2=cn2PT. Moreover, (5.4) holds (see Figure 1). It remains to verify the condition (5.17). The values of the coefficients for the equations of b0 and λ1 are given by the following formulas, obtained by using the equation verified by cn and the explicit expressions given by Lemma 4.8. Due to the complicated nature of the expressions, the dependence of E and K on k will be left implicit:
Therefore,
Thus, Proposition (5.4) applies, providing an unstable eigenvalue of J(Lcn)θ for θ=π2Kε, and all 0<εε0. It follows in particular that cn is unstable against perturbations with period 4nK, where n is the smallest even integer greater than πKε0. Indeed, let ε=πKn. Then θ=π2Kε=n22nπK, and by (5.3), we have σ(JLθP2K)σ(JLP4nK). This concludes the proof of Theorem 5.3.■

5.2 Numerical spectra

We have tested numerically the spectra of the different operators involved. To this aim, we used a fourth order centered finite difference discretization of the second derivative operator. Unless otherwise specified, we have used 210 grid points. The spectra are then obtained using the built in function of our scientific computing software (Scilab). Whenever the spectra can be theoretically described, the theoretical description and our numerical computations are in good agreement.

We start by the presentation of the spectra of JLpq, for pq=cn,dn,sn on P4K. 

Observation 5.6

On P4K, the spectrum of JLpq is such that

  • if pq=sn then σ(JLsn)iR for all k(0,1),

  • if pq=cn, then σ(JLcn)iR for all k(0,1), including when k>kc,

  • if pq=dn, then JLdn admits two double eigenvalues ±λ with λ>0 and the rest of the spectrum verifies (σ(JLdn){±λ})iR for all k(0,1).

The numerical observations for cn and dn at k=0.95 are represented in Figure 2.□

σ(JLcn) (left) and σ(JLdn) (right) on P4K for k = 0.95.
Fig. 2.

σ(JLcn) (left) and σ(JLdn) (right) on P4K for k = 0.95.

We then compare the results of Theorem 5.3 with the numerical results. In Figure 3, we have drawn the numerical spectrum of JLcn as an operator on L2(R). To this aim, we have used the Bloch decomposition of the spectrum of JLcn given in (5.2): we computed the spectrum of J(Lcn)θ for θ in a discretization of (0,π2K] and we have interpolated between the values obtained to get the curve in plain (blue) line. In order to keep the computation time reasonable, we have dropped the number of space points from 210 to 28. We then have drawn in dashed (red) the straight lines passing through the origin and the points whose coordinates are given in the complex plane by ±λ1,±λ¯1, λ1 given in the proof of Proposition 5.4. The picture shows that the dashed (red) line are tangent to the plain (blue) curve, thus confirming λ1 as the first order in the expansion for the eigenvalue emerging from 0 performed in Proposition 5.4.

σ(JLcn) on L2(R) for k = 0.9 (plain (blue) curve), first order asymptotic around 0 (dashed (red) lines).
Fig. 3.

σ(JLcn) on L2(R) for k = 0.9 (plain (blue) curve), first order asymptotic around 0 (dashed (red) lines).

Numerically, eigenvalues on the Number 8 curve in Figure 3 are simple, and move from the origin toward the intersection points of the Number 8 curve with the imaginary axis, when θ is decreased from π/(2K) to 0+.

These eigenvalues are simple because we did the Block decomposition (5.2) in P2K with θ[0,2π/T)=[0,π/K), and cn is only in P2K(1), not in P2K. Thus, it is in the kernel of L+θ only for θ=π/(2K). The bifurcation occurs only near θ=π/(2K), not at θ=0.

In contrast, Rowlands [30] did the Block decomposition in P4K with θ[0,π/(2K)). We have cnP4K, and cn is in the kernel of L+θ only for θ=0. The bifurcation occurs only near θ=0.

These two approaches are essentially the same, and our approach does not give a new instability branch.

6 Numerics

We describe here the numerical experiments performed to understand better the nature of the Jacobi elliptic functions as constrained minimizers of some functionals. To this aim, we use a normalized gradient flow approach related to the minimization problem (3.3).

6.1 Gradient flow with discrete normalization

It is relatively natural when dealing with constrained minimization problems like (3.3) and (3.4) to use the following construction. Define an increasing sequence of time 0=t0<<tn and take an initial data u0. Between each time step, let u(t,x) evolve along the gradient flow
At each time step tn, the function is renormalized so as to have the desired mass and momentum. The renormalization for the mass is obtained by a straightforward scaling:
(6.1)
When there is no momentum, like in the minimization problems (3.1), (3.4), and only real-valued functions are considered, such approach to compute the minimizers was developed by Bao and Du [4].

However, dealing with complex valued solutions and with an additional momentum constraint as in problems (3.3), (3.5) turns out to make the problem more challenging and to our knowledge little is known about the strategies that one can use to deal with this situation (see [9] for an approach on a related problem).

To construct un in such a way that (un)=p, a simple scaling is not possible for at least two reasons. First of all, if p=0, a scaling would obviously lead to failure of our strategy. Second, even if p0, as we are already using a scaling to get the correct mass, making a different scaling to obtain the momentum constraint will result into a modification of the mass. To overcome these difficulties, we propose the following approach.

Recall that, as noted in [4], the renormalizing step (6.1) is equivalent to solving exactly the following ordinary differential equation:
(6.2)
Inspired by this remark, we consider the following problem, which we see as the equivalent of (6.2) for the momentum renormalization:
(6.3)
where we want to choose the values of ϖn in such a way that (u(tn+1))=p. To this aim, we need to solve (6.3). Note that (6.3) is a partial differential equation, whereas (6.2) was only an ordinary differential equation. We make the following formal computations, which can be justified if the functions involved are regular enough. As we work with periodic functions, we consider the Fourier series representation of u, that is
with the Fourier coefficients
Then, (6.3) becomes
For each jZ and for any tn<t<tn+1, the solution is
and, therefore, the solution of (6.3) is
Using this Fourier series expansion of u, we have
We determine implicitly the value of ϖn, by requiring that ϖn is such that
In practice, it might not be so easy to compute ϖn and, therefore, we shall use the following approximation. We replace the exponential by its first-order Maclaurin polynomial. We get
Therefore, an approximation for ϖn is given by ϖ˜n, which is defined implicitly by
Solving for ϖ˜n, we obtain
We can further simplify the expression of ϖ˜n by remarking that
This gives
This is the value we will use in practice.

6.2 Discretization

Let us now further discretize our problem. We first present a semi-implicit time discretization, given by the following scheme:
where ϖ˜n is given by
and (cj(u˜n+1)) are the Fourier coefficients of u˜n+1. Note that the system is linear. 
Remark 6.1

If p=0, at the end of each step, un+1 has the desired mass and momentum. If p0, then un+1 only has the desired mass and it is unclear if the algorithm will still give convergence toward the desired mass–momentum constraint minimizer. We plan to investigate this question in further works.□

Finally, we present the fully discretized problem. We discretize the space interval T2,T2 by setting
We denote by unl the numerical approximation of u(tn,xl). Using the (backward Euler) semi-implicit scheme for time discretization and second-order centered finite difference for spatial derivatives, we obtain the following scheme:
(6.4)
(6.5)
(6.6)
where cj(u˜n+1)=1L+1l=0Lu˜n+1lei2πLjlδx.

As the system (6.4) is linear, we can solve it using a Thomas algorithm for tridiagonal matrix modified to take into account the periodic boundary conditions. The discrete Fourier transform and its inverse are computed using the built in Fast Fourier Transform algorithm.

We have not gone further in the analysis of the scheme presented above. As shown in the next section, the outcome of the numerical experiments is in good agreement with the theoretical results. We plan to further analyze and generalize our approach in future works.

7 Numerical solutions of minimization problems

Before presenting the numerical experiments, we introduce some notation for particular plane waves. Define

In the numerical experiments, we have chosen to fix k = 0.9. The period will be either T=2K(k) or T=4K(k). We use 212 grid points for the interval [T2,T2]. The time step will be set to 1. We decided to run the algorithm until a maximal difference of 10−3 between the absolute values of the moduli of ujl and the expected minimizer has been reached.

We made the tests with the following initial data:
(7.1)
Depending on the expected profile, we may have shifted uj so that a minimum or a maximum of its modulus is at the boundary. Since the problem is translation invariant, this causes no loss of generality.

Since the initial data u0 in (7.1) do not match the required mass/momentum, u1 are very different from u0. Thus, (7.1) is a random choice, and this shows up in the rapid drop from t0 to t1 in Figure 4. The idea is to show that the choice of initial data is not important for the algorithm and that no matter from where the algorithm is starting, it converges to the supposed minimizer (unless the initial data have some symmetry preserved by the algorithm).

For m=π28K<π2bT, focusing, periodic case.
Fig. 4.

For m=π28K<π2bT, focusing, periodic case.

7.1 Minimization among periodic functions

Minimization among periodic functions is completely covered by the theoretical results’ Propositions 3.2 and 3.3. We have performed different tests using the scheme described in (6.4) and (6.6) and we have found that the numerical results are in good agreement with the theoretical ones.

7.1.1 The focusing case

In all the experiments performed in this case, we have tested the scheme with and without the momentum renormalization step (6.6) and we have obtained the same result each time. This confirms that in the periodic case the momentum constraint plays no role (see (i) in Proposition 3.2, and Proposition 3.3). In what follows, we present only the results obtained using the full scheme with renormalization of mass and momentum.

We fix T=2K(k) and b=2. We first perform an experiment to verify the agreement with case (ii) in Proposition 3.2. Let m=π28K<π2bT. With each initial data in (7.1), we observe convergence towards the constant solution, hereby confirming case (ii) of Proposition 3.2. The results are presented in Figure 4 for initial data (c) of (7.1). The requested precision is achieved after 14 time steps.

The second experiment that we perform is aimed at testing case (iv) of Proposition 3.2. Let m=M(dn)=E(k). Once again we observe a good agreement between the theoretical prediction and the numerical experiment. The results are presented in Figure 5 for initial data (c) of (7.1). The requested precision is achieved after 16 time steps.

For m=M(dn)=E(k), focusing, periodic case.
Fig. 5.

For m=M(dn)=E(k), focusing, periodic case.

All the other experiments that we have performed show a good agreement with the theoretical results in the focusing case for minimization among periodic functions. To avoid repetition, we give no further details here.

7.1.2 The defocusing case

We now present the experiment in the defocusing case. We have used b=2k2 and T=4K. We have tested the algorithm with and without the momentum renormalization step (6.6), obtaining the same results. The results are presented in Figure 6 for initial data (c) of (7.1) and mass constraint m=M(sn)=2(KE)k2. The requested precision is achieved after seven time steps.

For m=M(sn)=2(K−E)k2, defocusing, periodic case.
Fig. 6.

For m=M(sn)=2(KE)k2, defocusing, periodic case.

7.2 Minimization among half-anti-periodic functions

We will in that case add an additional step in the algorithm in which we keep only the anti-periodic part of the function. This way it will not matter whether or not our initial data has the right anti-periodicity, since anti-periodicity will be forced at each iteration of the algorithm.

7.2.1 The focusing case

We compare in this section the numerical results with Proposition 3.4. We have used b=2k2 and T=4K. The tests performed show a good agreement between the numerics and the theoretical result. We present in Figure 7 the result for initial data (c) of (7.1) and mass constraint m=M(cn)=2(E(1k2)K)/k2

For m=M(cn)=2(E−(1−k2)K)/k2, focusing, anti-periodic case.
Fig. 7.

For m=M(cn)=2(E(1k2)K)/k2, focusing, anti-periodic case.

7.2.2 The defocusing case

We finally turn out to the defocusing case, still imposing anti-periodicity. We have used b=2k2 and T=4K.

We have tested the algorithm without the momentum renormalization step (6.6) and confirmed the theoretical result Proposition 3.6, which states that a plane wave is the minimizer. We present the result in Figure 8 for initial data (c) of (7.1) and mass constraint m=M(sn)=2(KE)k2. Note a plateau in the two graphs of Figure 8. This is due to the fact that the sequence remains for some time close to sn (which is the expected minimizer if we impose in addition the momentum constraint), before eventually converging to the plane wave minimizer.

For m=M(sn)=2(E(k)−K)/k2, defocusing, anti-periodic case without momentum constraint.
Fig. 8.

For m=M(sn)=2(E(k)K)/k2, defocusing, anti-periodic case without momentum constraint.

Finally, we run the full algorithm with mass and momentum renormalization for mass constraint m=M(sn)=2(KE)k2 and 0 momentum constraint. No theoretical result is available in this case. We made the following observation, which confirms Conjecture 3.7. 

Observation 7.1

The function sn is a minimizer for problem (3.5) with m=M(sn).□

We present in Figure 9 the result of the experiment with full algorithm for initial data (c) of (7.1) and mass constraint m=M(sn)=2(KE)k2.

For m=M(sn)=2(E(k)−K)/k2, defocusing, anti-periodic case with momentum constraint.
Fig. 9.

For m=M(sn)=2(E(k)K)/k2, defocusing, anti-periodic case with momentum constraint.

Acknowledgments

We are grateful to Bernard Deconinck and Dmitri Pelinovsky for useful remarks on a preliminary version of this paper.

Funding

The work of S.G. is partially supported by NSERC Grant 251124-12. The work of S.L.C. is partially supported by ANR-11-LABX-0040-CIMI within the program ANR-11-IDEX-0002-02 and ANR-14-CE25-0009-01. The work of T.T. is partially supported by NSERC Grant 261356-13.

References

1

Abramowitz
,
M.
and
I. A.
Stegun
Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables, volume 55 of National Bureau of Standards Applied Mathematics Series
Washington, D.C.
:
U.S. Government Printing Office
,
1964
.

2

Angulo Pava
,
J.
Nonlinear stability of periodic traveling wave solutions to the Schrödinger and the modified Korteweg–de Vries equations
.”
Journal of Differential Equations
235
, no.
1
(
2007
):
1
30
.

3

Angulo Pava
,
J.
Nonlinear Dispersive Equations: Existence and Stability of Solitary and Periodic Travelling Wave Solutions, volume 156 of Mathematical Surveys and Monographs.
Providence, RI
:
American Mathematical Society
,
2009
.

4

Bao
,
W.
and
Q.
Du
. “
Computing the ground state solution of Bose–Einstein condensates by a normalized gradient flow
.”
SIAM Journal of Scientific Computing
25
, no.
5
(
2004
):
1674
97
.

5

Bottman
,
N.
,
B.
Deconinck
and
M.
Nivala
. “
Elliptic solutions of the defocusing NLS equation are stable
.”
Journal of Physics A
44
, no.
28
(24,
2011
):
285201
.

6

J.
Bourgain
.
Global Solutions of Nonlinear Schrödinger Equations, volume 46 of American Mathematical Society Colloquium Publications
.
American Mathematical Society
,
Providence, RI
,
1999
.

7

Cazenave
,
T.
Semilinear Schrödinger Equations
New York
:
New York University—Courant Institute
,
2003
.

8

Cazenave
,
T.
and
P.-L.
Lions
. “
Orbital stability of standing waves for some nonlinear Schrödinger equations
.”
Communications in Mathematical Physics
85
, no.
4
(
1982
):
549
61
.

9

Chiron
,
D.
and
C.
Scheid
. “
Travelling waves for the Nonlinear Schrödinger Equation with general nonlinearity in dimension two
.”
Journal of Nonlinear Science
26
, no.
1
(
2016
):
171
231
.

10

Eastham
,
M. S. P.
The Spectral Theory of Periodic Differential Equations. Texts in Mathematics (Edinburgh)
New York
:
Scottish Academic Press, Edinburgh; Hafner Press
,
1973
.

11

Fibich
,
G.
The Nonlinear Schrödinger Equation, volume 192 of Applied Mathematical Sciences
Cham
:
Springer
,
2015
.

12

Gallay
T.
Existence et stabilité des fronts dans l'équation de Ginzburg–Landau à une dimension. PhD thesis, Université de Genève,
1994
.

13

Gallay
,
T.
and
M.
Hărăgus
. “
Orbital stability of periodic waves for the nonlinear Schrödinger equation
.”
Journal of Dynamics and Differential Equations
19
, no.
4
(
2007
):
825
65
.

14

Gallay
,
T.
and
M.
Hărăguş
. “
Stability of small periodic waves for the nonlinear Schrödinger equation
.”
Journal of Differential Equations
234
, no.
2
(
2007
):
544
81
.

15

Gallay
,
T.
and
D.
Pelinovsky
. “
Orbital stability in the cubic defocusing NLS equation: I. Cnoidal periodic waves
.”
Journal of Differential Equations
258
, no.
10
(
2015
):
3607
38
.

16

Geyer
,
A.
and
D. E.
Pelinovsky
. “
Spectral stability of periodic waves in the generalized reduced ostrovsky equation
.”
Letters in Mathematical Physics
(
2017
):
1
22
. doi:10.1007/s11005-017-0941-3.

17

Gradshteyn
,
I. S.
and
I. M.
Ryzhik
Table of Integrals, Series, and Products
, eighth edition.
Amsterdam
:
Elsevier/Academic Press
,
2015
.

18

Grillakis
,
M.
,
J.
Shatah
and
W. A.
Strauss
. “
Stability theory of solitary waves in the presence of symmetry I
.”
Journal of Functional Analysis
74
, no.
1
(
1987
):
160
97
.

19

Grillakis
,
M.
,
J.
Shatah
and
W. A.
Strauss
. “
Stability theory of solitary waves in the presence of symmetry II
.”
Journal of Functional Analysis
94
, no.
2
(
1990
):
308
48
.

20

Hărăguş
,
M.
Kapitula
,
T.
. “
On the spectra of periodic waves for infinite-dimensional Hamiltonian systems
.”
Physica D
237
, no.
20
(
2008
):
2649
71
.

21

Ivey
,
T.
and
S.
Lafortune
. “
Spectral stability analysis for periodic traveling wave solutions of NLS and CGL perturbations
.”
Physica D
237
, no.
13
(
2008
):
1750
72
.

22

Kapitula
,
T.
,
P. G.
Kevrekidis
and
B.
Sandstede
. “
Counting eigenvalues via the Krein signature in infinite-dimensional Hamiltonian systems
.”
Physica D
195
, no.
3–4
(
2004
):
263
82
.

23

Kapitula
,
T.
,
P. G.
Kevrekidis
and
B.
Sandstede
. “
Addendum: “Counting eigenvalues via the Krein signature in infinite-dimensional Hamiltonian systems” [Physics D 195, no. 3–4 (2004), 263–282; mr2089513]
.”
Physica D
201
, no.
1–2
(
2005
):
199
201
.

24

Lawden
,
D. F.
Elliptic Functions and Applications, volume 80 of Applied Mathematical Sciences
New York
:
Springer-Verlag
,
1989
.

25

Le Coz
,
S.
,
D.
Li
and
T.-P.
Tsai
. “
Fast-moving finite and infinite trains of solitons for nonlinear Schrödinger equations
.”
Proceedings of the Royal Society of Edinburgh Section A
145
, no.
6
(
2015
):
1251
82
.

26

Le Coz
,
S.
and
T.-P.
Tsai
. “
Infinite soliton and kink-soliton trains for nonlinear Schrödinger equations
.”
Nonlinearity
27
, no.
11
(
2014
):
2689
2709
.

27

Ma
,
Y. C.
and
M. J.
Ablowitz
. “
The periodic cubic Schrödinger equation
.”
Studies in Applied Mathematics
65
, no.
2
(
1981
):
113
58
.

28

Marzuola
,
J. L.
and
D. E.
Pelinovsky
. “
Ground state on the dumbbell graph
.”
Applied Mathematics Research eXpress
2016
, no.
1
(
2016
):
98
145
.

29

Reed
,
M.
and
B.
Simon
Methods of Modern Mathematical Physics. IV. Analysis of Operators
New York-London
:
Academic Press [Harcourt Brace Jovanovich, Publishers]
,
1978
.

30

Rowlands
,
G
. “
On the stability of solutions of the non-linear Schrödinger equation
.”
IMA Journal of Applied Mathematics
13
, no.
3
(
1974
):
367
77
.

31

Sulem
,
C.
and
P.-L.
Sulem
The Nonlinear Schrödinger Equation, volume 139 of Applied Mathematical Sciences
New York
:
Springer-Verlag
,
1999
) Self-Focusing and Wave Collapse.