Home Low regularity conservation laws for Fokas-Lenells equation and Camassa-Holm equation
Article Open Access

Low regularity conservation laws for Fokas-Lenells equation and Camassa-Holm equation

  • Minjie Shan , Mingjuan Chen EMAIL logo , Yufeng Lu and Jing Wang
Published/Copyright: June 8, 2024

Abstract

In this article, we mainly prove low regularity conservation laws for the Fokas-Lenells equation in Besov spaces with small initial data both on the line and on the circle. We develop a new technique in Fourier analysis and complex analysis to obtain the a priori estimates. It is based on the perturbation determinant associated with the Lax pair introduced by Killip, Vişan, and Zhang for completely integrable dispersive partial differential equations. Additionally, we also utilize the perturbation determinant to derive the global a priori estimates for the Schwartz solutions to the Camassa-Holm (CH) equation in H 1 . Even though the energy conservation law of the CH equation is a fact known to all, the perturbation determinant method indicates that we cannot get any conserved quantities for the CH equation in H k except k = 1 .

MSC 2010: 37K10; 35Q55

1 Introduction

The method introduced by Killip et al. [43], which was based on the power series expansion of the perturbation determinant associated with the Lax pair, provides a powerful technique for obtaining the low regularity conservation laws for nonlinear integrable partial differential equations (PDEs). By utilizing this method, we develop a new technique in Fourier analysis and complex analysis to study the Fokas-Lenells (FL) equation

(1.1) i u t ν u t x + γ u x x + σ u 2 ( u + i ν u x ) = 0 ,

where u : × R C ( = R or T = R Z ), γ and ν are two real parameters, and σ = ± 1 . This equation was first derived by Fokas applying the bi-Hamiltonian method in [28]. In optics, considering higher order linear and nonlinear optical effects, the FL equation is used as a model to describe the femtosecond pulse propagation through single mode optical silica fiber [50].

Incidentally, the following Camassa-Holm (CH) equation on R or T is also considered:

(1.2) u t u t x x + 2 ω u x + 3 u u x = 2 u x u x x + u u x x x ,

where u : × R R and ω is a positive constant. The CH equation describes unidirectional propagation of wave in shallow water. It is derived physically by Camassa and Holm [7] by approximating the Hamiltonian for Euler’s equations in the shallow water regime. In fact, this equation first appeared in the paper of Fuchssteiner and Fokas [30] on hereditary symmetries, they proved that (1.2) is formally integrable as a bi-Hamiltonian generalization of KdV. But it came to be remarkable in the work of Camassa and Holm [7] where the peakon was described. The CH equation also arises in the study of axially symmetric waves motion in hyperelastic rods [21,23]. Its high-frequency limit models nematic liquid crystals [4,39].

Integrable equations enjoy several special properties including the existence of a Lax pair, infinitely many conserved quantities, as well as a bi-Hamiltonian formulation. Actually, the last property provides an important algorithmic approach of constructing integrable equations. Suppose that for all values of the parameter λ the operator θ 1 + λ θ 2 is Hamiltonian, then

(1.3) q t = θ 2 θ 1 1 q x

gives an integrable equation. If we define

(1.4) θ 1 = x , θ 2 = α x + β x 3 + γ ( q x + x q ) , α , β are constants ,

when α = 0 , β = 1 , γ = 2 , then (1.3) yields the KdV equation

q t + q x x x + 6 q q x = 0 .

If we define q = u u x x , and replace θ 1 = x by θ 1 = x x 3 , then (1.3) with α = 2 ω , β = 0 , γ = 1 yields the CH equation (1.2).

If we define r = σ q ¯ , and the compatible pair of Hamiltonian operators

σ 3 = 1 0 0 1 and θ 2 = γ x + q x 1 r q x 1 q r x 1 r γ x + r x 1 q ,

then the system

q r t = i θ 2 σ 3 q r x

reduces to the cubic nonlinear Schrödinger (NLS) equation

(1.5) i q t + γ q x x + σ q 2 q = 0 .

Furthermore, if we define q = u + i ν u x , and replace σ 3 by θ 1 = σ 3 + i ν I x , where I is the 2 × 2 identity matrix, then the system

q r t = i θ 2 θ 1 1 q r x

gives the FL equation (1.1). Therefore, the CH equation is considered as an integrable generalization of the KdV equation, and the FL equation is an integrable generalization of the NLS equation.

Utilizing the bi-Hamiltonian structure, the first few conservation laws and the Lax pair of the FL system were constructed by Lenells and Fokas [53]. Moreover, the Lax pair was used to solve the initial value problem on the line with inverse scattering methods under the assumption that the initial data have sufficient smoothness and decay. We may assume that ν has the same sign as γ in equation (1.1), otherwise we just replace u ( x , t ) with u ( x , t ) . Let β = γ ν > 0 and μ = 1 ν . Then, the transformation

u μ β e i μ x q , σ σ ,

converts (1.1) into

(1.6) q t x + β μ 2 q 2 i β μ q x β q x x + i σ β μ 2 q 2 q x = 0 .

The method in our work is applicable to both σ = ± 1 , and the procedures are the same except for the signs in certain places, so we only consider the case σ = 1 . Furthermore, we take β = μ = 1 for simplicity, hence (1.6) can be written as

(1.7) q t x + q 2 i q x q x x + i q 2 q x = 0 .

There are infinitely many conservation laws for the FL flow (1.7) and the first three (see [53] Section 2) are

(1.8) H 0 = q x 2 d x , H 1 = i q x ( q ¯ + i q ¯ x ) d x = i q ¯ x ( q i q x ) d x , H 2 = q ¯ ( q + 2 i q x + q x x ) + i 2 q q x q ¯ 2 d x .

(1.7) admits a Lax pair with operator

(1.9) L ( t ; q ) = + i κ 2 κ q x κ q ¯ x i κ 2 .

The FL equation is also closely related to the derivative nonlinear Schrödinger (DNLS) equation, i.e., the nonlinearity q 2 q in (1.5) is replaced by ( q 2 q ) x . It is integrable and belongs to the DNLS hierarchy [53], which is related to the Kaup-Newell spectral problem. In fact, it is the first negative flow of the integrable hierarchy of the DNLS equation [50]. Actually, we cannote that the operator L ( t ; q ) in (1.9) has the same form as the x -part of the Lax pair for the DNLS equation derived in the pioneering work of Kaup and Newell [41].

The DNLS equation admits the scaling symmetry and L 2 is the scaling critical space. It is worth noting that the best global well-posedness in L 2 ( R ) for this equation has been obtained by Harrop-Griffiths et al. [34] by the second-generation method of commuting flows. Before this work, multiple authors sought to obtain well-posedness in H s for s as small as possible. We now review some but not exhaustive results. Local well-posedness in H 1 2 was obtained by Takaoka [67] on the real line and Herr [36] on the circle by fixed point arguments. Biagioni and Linares [3] showed that the data-to-solution mapping fails to be locally uniformly continuous on the real line below H 1 2 . This indicates that s = 1 2 is the optimal index of proving local well-posedness via the contraction mapping principle. By employing exact or approximate conservation laws, one can extend local-in-time solutions to global solutions. However, the conservation laws (including mass, Hamiltonian, etc.) for the DNLS equation are not coercive for large initial data, specifically, when q 0 L 2 is large. Thus, the subsequent global results were obtained under the restriction of bounded initial data q 0 L 2 . Wu [71,72] proved the global well-posedness in energy space with the initial data q 0 H 1 ( R ) and q 0 L 2 < 2 π , and Mosincat and Oh got a similar result on the torus [60]. Subsequently, Guo and Wu [33] proved global well-posedness in H 1 2 ( R ) for q 0 L 2 < 2 π by I-method, and Mosincat [59] showed global well-posedness in the circle case under the same L 2 -smallness condition. Recently, by combining the known well-posedness theory with an in-depth analysis of the transmission coefficient, Bahouri and Perelman [2] got rid of the L 2 -smallness condition and proved that the DNLS equation is globally well-posed in H 1 2 ( R ) . For the lower regularity, Guo [31] established a priori estimates for s > 1 4 on the real line by short-time Fourier restriction method, and Schippa [64] extended this result to periodic boundary conditions. Recently, Klaus and Schippa [45] established low regularity a priori estimates for the DNLS equation in the full subcritical spaces B r , 2 s ( 0 < s < 1 2 ) with small q 0 L 2 . The microscopic conservation laws for the Schwartz solutions of the DNLS equation with small mass were also worked out by Tang and Xu [69].

The FL equation has been studied in various interesting manners. For example, solutions of the FL equation can be derived by means of the Riemann-Hilbert method [53], inverse scattering transform [80], dressing chain [51], algebrogeometric method [81], Darboux transformation [35,78], and a variable separation technique [70]. The multisoliton solutions including dark solitons and bright solitons were obtained in [51,54,55]. Fan and his co-authors in [12,75] studied the long-time asymptotics for the FL equation with decaying initial value problem by means of the Riemann-Hilbert approach. The initial-boundary value problems of the FL equation have been studied in [52,73].

Less is known about the well-posedness theory for the FL equation. Fokas and Himonas [29] obtained that the local well-posedness of the periodic initial value problem for the FL equation (1.1) in H s ( T ) for s > 3 2 , if 1 ν is not an integer. When 1 ν is an integer, they gave the local well-posedness for the nonlocal version of the FL equation in the corresponding homogeneous Sobolev spaces. Cheng and Fan [11] used inverse scattering method based on the Riemann-Hilbert problem to obtain the existence of global solutions to the Cauchy problem for the FL equation on the line without the small-norm assumption on initial data u 0 H 3 ( R ) H 2 , 1 ( R ) . Recently, they used a newly modified Darboux transformation to obtain the global well-posedness result in H 3 ( R ) H 2 , 1 ( R ) when the initial data include solitons [13]. However, there are major technical difficulties to use inverse scattering techniques in unweighted Sobolev spaces. For instance, on the line, the decay of the data is insufficient when inverse scattering method is applied.

The principal goal of this work is to obtain the low regularity conservation laws for the FL equation in the Besov spaces. For the NLS equation on the line, Koch and Tataru [46] used the transmission coefficient to obtain almost conserved H s -energies for all s > 1 2 . Killip et al. [43] developed a method to obtain conservation laws both on the real line R and on the circle T for a general class of completely integrable dispersive equations, such as KdV, cubic NLS and mKdV (which lie within the famous AKNS/ZS framework). Then, they showed the global well-posedness for the KdV equation in H 1 by using these conservation laws [44]. Talbut [68] used the same approach to obtain conservation laws at negative regularity ( 1 2 < s < 0 ) for the Benjamin-Ono (BO) equation. Killip et al. [42] showed very recently that the BO equation is well-posed, both on the line and on the circle, in the Sobolev spaces H s for s > 1 2 by using a new gauge transformation and a modified Lax pair representation of the full hierarchy. As mentioned above, the conservation laws for the DNLS equation were studied in [45,69].

Motivated by these work, we will show that the determinant

log det ( + i κ 2 ) 1 0 0 ( i κ 2 ) 1 + i κ 2 κ q x κ q ¯ x i κ 2

given by

(1.10) α FL ( κ ˜ ; q ( t ) ) Re = 1 i 1 κ ˜ tr { [ ( κ ˜ ) 1 q x ( κ ˜ + ) 1 q ¯ x ] } ,

where κ ˜ = i κ 2 (the tilde may be dropped later on), is conserved for solutions of (1.7). Then, by using this conserved quantity, we obtain the growth of Besov norms of the solution. Now, we state our main result.

Theorem 1.1

Let 0 < s < 1 2 and 1 r . Assume that q ( t , x ) is a Schwartz solution to (1.7) on R or R Z with small initial data condition q x ( 0 ) B 2 , r s c 1 , then we have

(1.11) q x ( t ) B 2 , r s q x ( 0 ) B 2 , r s .

Remark 1.2

Recalling Cheng and Fan’s global well-posedness results [11,13] in H 3 ( R ) H 2 , 1 ( R ) and Fokas and Himonas’s local well-posedness results [29] in H ˙ s ( T ) ( s > 3 2 ), one may ask whether (1.11) holds true for s 1 2 ? This question is of fundamental importance, since if the answer is positive, by employing the conservation laws, we can extend local well-posedness to global well-posedness. Unfortunately, we are currently unable to prove (1.11) for s 1 2 . In this work, the condition s < 1 2 is necessary. Here is the key point. The remainder of the series expansion of the perturbation determinant for the FL equation is controlled by

= 2 tr { [ κ R q x R + q ¯ x ] } κ 4 s q x ( t ) H s 4 ,

if 0 < s < 1 4 and κ 1 , see (5.6). Note that Z κ 0 defined in (5.8) as an equivalent norm of B 2 , r s is written as the summation of dyadic block N . Taking κ = κ 0 N , then

N 2 N N r s ( κ 0 N ) 2 s r <

yields that s < 2 s . Hence, we need s < 1 2 . One can see (5.10) for more details.

Remark 1.3

The small initial data assumption is essential. To guarantee convergence and term-by-term differentiation of the series (1.10), we need the smallness condition

tr { κ ( κ ) 1 q x ( κ + ) 1 q ¯ x } c < 1 .

For the NLS equation, one can choose sufficiently large κ such that [43]

tr { ( κ ) 1 q ( κ + ) 1 q ¯ } log 4 + ξ 2 κ 2 q ˆ ( ξ ) 2 4 κ 2 + ξ 2 d ξ c < 1 .

However, for the FL equation, there is an additional κ in each term. This directly leads to that the power of κ appeared in the estimate of Besov norm of the solution is positive rather than negative, see (5.12). Therefore, we cannot obtain smallness like the NLS case, in which the convergence of the series is acquired by choosing κ sufficiently large. Because of the same reason, Theorem 1.1 cannot be extended to the negative regularity.

Local well-posedness of the CH equation in Sobolev spaces and Besov spaces has been studied [17,24,25,47,48]. It was shown that the CH equation is well-posed in H s for s > 3 2 and ill-posed in H s for s < 3 2 in the sense that the data-to-solution map fails to be uniformly continuous. In the critical Sobolev space H 3 2 , there exists norm inflation and ill-posedness [32]. Xin and Zhang [74] proved the existence of a global-in-time weak solution to the Cauchy problem of the CH equation with initial data in H 1 ( R ) . We refer to [5,19,38] on existence and uniqueness of global weak solutions to the CH equation. Xu and Yang [76] established the local well-posedness of solutions with the low regularity in the periodic setting for a class of one-dimensional generalized Rotation-Camassa-Holm equation.

The inverse scattering transform method is used to study the initial value problem for the CH equation in [15,18,40,49]. Deift and Zhou [26,27] developed a very effective tool called the nonlinear steepest descent method to obtain the long-time asymptotic behavior of systems, see also [57,75].

There are several important features for the CH equation. For example, (1.2) is an infinite-dimensional Hamiltonian system which is completely integrable [14,15]. The CH equation possesses non-smooth peakon-type solitary or periodic traveling wave solutions, in sharp contrast to the solitary waves for the KdV equation, which are always analytic [1,6,8,16]. A peakon is a weak solution in some Sobolev space with corner at its crest. Constantin further showed the stability of peakons and orbital stability of solitary wave solution for the CH equation in [20,22]. Stability of peakons for modified CH equation was obtained in [62]. Moon [58] obtained the existence of peaked traveling wave solution of μ -Novikov equation which is a linear combination of the Novikov equation and CH equation. Xu and his collaborators [10,77] proved local well-posedness, rigidity property, and asymptotic stability of peakons in the energy space H 1 ( R ) for some generalized shallow water equations.

An infinite number of conservation laws of the KdV equation help in deriving the necessary a priori estimates for solutions in H k ( R ) for every k Z + . However, although (1.2) is integrable and has infinitely many conservation laws, global existence of regular solutions is guaranteed only for a special class of initial data, and only the H 1 ( R ) norm is preserved for smooth solutions to (1.2). In fact, it has been observed by Camassa and Holm [7], Constantin and Escher [17], and McKean [56] that finite-time blow-up of smooth solutions always occurs for a large class of initial data. In particular, McKean [56] gave a necessary and sufficient criterion on the initial data for the finite-time formulation of singularities in a smooth solution to the CH equation.

Without loss of generality, we fix ω = 1 in the CH equation (1.2) and assume that u ( x ) and u 0 ( x ) are in Schwartz space. The Lax pair formulation transcends the distinction between the problem on the line with decaying and periodic data. (1.2) admits the following Lax pair:

(1.12) Ψ x = M Ψ , Ψ t = N Ψ ,

where

M = 0 1 1 4 + λ ( q + 1 ) 0 ,

N = 1 2 u x 1 2 λ u ( 1 2 λ u ) 1 4 + λ ( q + 1 ) + 1 2 u x x 1 2 u x ,

where q = u u x x and λ is a spectral parameter.

We consider the perturbation determinant

(1.13) det x 1 1 4 + λ x 1 x 1 1 4 + λ ( q + 1 ) x ,

which formally represents the ratio of the “characteristic polynomials” of Lax operator and that of the Schrödinger operators with no potential. Perturbation determinant plays an important role in perturbation and scattering theory [63,65,79]. After simple calculation, one can see that the determinant of (1.13) can be rewritten as

det I + λ R 0 q 0 λ R 0 x q 0 det ( I + A ) ,

where R 0 = ( x 2 + κ 2 ) 1 and κ 2 = 1 4 + λ . R 0 is well-defined whenever κ > 0 . Then

ln det ( I + A ) = tr ln ( I + A ) = = 1 ( 1 ) 1 tr { A } = = 1 ( 1 ) 1 tr { ( λ R 0 q ) } .

The leading term in the Fredholm expansion is

tr ( λ R 0 q ) = λ 2 κ R q ( x ) d x = λ 2 κ R ( u u x x ) d x ,

which is actually a conservation quantity of the CH equation (1.2). Obviously, this will not extend to functions u in H 1 . By using a device of Hilbert [37], we consider the regularized 2-determinant to drop the = 1 term

(1.14) det 2 ( I + A ) det ( I + A ) exp { tr ( A ) } .

Furthermore, since R 0 q R 0 is self-adjoint, we consider the following series:

(1.15) α CH ( κ ; q ( t ) ) = ln ( det 2 ( I + A ) ) = = 2 + ( 1 ) tr { ( λ R 0 q R 0 ) } .

Now, this renormalization allows us to extend the notion of determinant to the class of Hilbert-Schmidt (HS) operators. By a simple computation, we know that the operator R 0 q R 0 is HS if and only if q H 1 . Precisely, the HS norm of R 0 q R 0 is comparable to u H 1 (see Corollary 3.3). Then, we use the conservation of the perturbation determinant (1.15) associated with the Lax pair to show the following a priori estimates for Schwartz solution to (1.2).

Theorem 1.4

Assume that u is a Schwartz solution to (1.2) on R or R Z . Then, we have

(1.16) 1 10 u ( 0 ) H 1 < u ( t ) H 1 < 10 u ( 0 ) H 1 .

Remark 1.5

In order to ensure the convergence of (1.15), we need the smallness of λ , which implies that κ in R 0 is located around 1/2. Unlike the KdV and NLS equation dealt with in [43], we cannot utilize the parameter κ to establish the relationship between the H s -norm and the principal term in series expansion of the perturbation determinant for the CH equation. Therefore, apart from H 1 , we cannot obtain any conserved quantities of the CH equation for other regularity indices.

There are several new ingredients in this study. First, the method introduced by Killip, Vişan, and Zhang for completely integrable dispersive PDEs is applied to general completely integrable equations. These equations, like KdV, cubic NLS, DNLS, and BO, enjoy the invariant scaling symmetry, but the FL equation and the CH equation do not. Second, we use the Fourier transform to compute the trace of operators in the Schatten class (Lemma 2.1), and develop a new approach in Fourier analysis and complex analysis to estimate the leading term of α ( κ ; q ( t ) ) . This method is not limited to the explicit kernel of the operators, but also applicable to general kernels, see the proof of Proposition 3.1 and Proposition 3.2 for more details. Finally, the proof of conservation of perturbation determinant for the FL equation is non-trivial. In order to prove Theorem 4.2, the key point is to prove

tr { ( R q x R + q ¯ x ) 1 ( R q 2 q x R + q ¯ x R q x R + q 2 q ¯ x ) } = κ tr { ( R q x R + q ¯ x ) ( R q R + q ¯ x + R q x R + q ¯ ) } for 1

where R = ( κ ) 1 and R + = ( + κ ) 1 . Formally, the term ( R q x R + q ¯ x ) R q R + q ¯ x in the right-hand side is comparable to R q 2 q R + q ¯ x because R and R + can be treated as integral operators. However, the term in the left-hand side is R q 2 q x R + q ¯ x . They differ by one-order derivative. In fact, we utilize the symmetry of operators’ trace to obtain some cancellation.

Organization of the paper. In Section 2, we present the notations and recall some basic facts about operator traces. In Section 3, we estimate the leading term of the perturbation determinants α ( κ ; q ( t ) ) associated with the FL and CH equations. In Section 4, we prove that the perturbation determinants α ( κ ; q ( t ) ) are conserved. Finally, by utilizing the conservation laws, we obtain the a priori estimates and complete the proof of Theorems 1.1 and 1.4 in Section 5.

2 Notations and preliminaries

In this section, we present the notations and several results that play a major role in the following estimates. For a detailed presentation, one can also refer to [43,45,65,66].

The Fourier transform is defined as

f ˆ ( ξ ) = 1 2 π R f ( x ) e i x ξ d x , ( F 1 f ) ( x ) = 1 2 π R f ( ξ ) e i x ξ d ξ

for functions on the line and

f ˆ ( ξ ) = 0 1 f ( x ) e i x ξ d x , ( F 1 f ) ( x ) = ξ 2 π Z f ( ξ ) e i x ξ

for functions on the circle T = R Z .

For s R , 1 r , we define Sobolev norms and Besov norms by

f H s ( R ) 2 = R ( 1 + ξ 2 ) s f ˆ ( ξ ) 2 d ξ , f H s ( R Z ) 2 = ξ 2 π Z ( 1 + ξ 2 ) s f ˆ ( ξ ) 2 , f B 2 , r s = f ˆ L 2 ( ξ 1 ) r + N 2 N N r s f ˆ L 2 ( N < ξ 2 N ) r 1 r ,

and with the usual interpretation when r = .

We recall some basic facts about operator traces and perturbation determinant. Denote I p the Schatten class of compact operators on L 2 with p -summable singular values. I p is complete and an embedded subalgebra of bounded operators. One can easily obtain the embedding I p I q from p q for p < q . I is the space of compact operators. I 1 is the space with elements of trace-class. If A is an operator on L 2 with continuous integral kernel K ( x , y ) , then the trace of operator A is defined by

tr ( A ) K ( x , x ) d x .

We refer to elements of I 2 by HS operators. An operator A on L 2 is HS class ( I 2 ) if and only if it admits an integral kernel K ( x , y ) L 2 , and define

A I 2 2 tr ( A A * ) = K ( x , y ) 2 d x d y .

Products of two or more HS operators A j with kernels K j ( x , y ) are of trace class, and the operator trace can be computed as

tr ( A 1 A n ) = K 1 ( x 1 , x 2 ) K 2 ( x 2 , x 3 ) K n ( x n , x 1 ) d x n d x 1 .

By Fubini’s theorem, this allows us to cycle the trace

tr ( A 1 A 2 A n ) = tr ( A 2 A n A 1 ) .

Specially,

tr ( A 2 ) = K ( x , y ) K ( y , x ) d y d x .

Moreover, if A is a trace class operator and B is a bounded operator on L 2 , then A B and B A are of trace class, and

tr ( A B ) = tr ( B A ) .

The class I p is two-sided ideals in the algebra of bounded linear operators on L 2 . Furthermore, we have

A B I p + B A I p A I p B op ,

and the Hölder-like estimate

A B I p A I q B I r

provided that 1 p = 1 q + 1 r , 1 p , q , r . For more details on operator theory and trace ideals, we refer to [65],[66, Chapter 3].

Next lemma tells us that one can consider the trace of operator from the Fourier transform point of view.

Lemma 2.1

(See Lemma 2.1 in [9]) Suppose that the operator A is given on the Fourier side by

A f ^ ( ξ ) = R m ( ξ , η ) f ˆ ( η ) d η ,

then the following results hold

(2.1) A * f ^ ( ξ ) = R m ( η , ξ ) ¯ f ˆ ( η ) d η ,

(2.2) tr ( A ) = 1 2 π R m ( ξ , ξ ) d ξ , A I 2 2 = 1 2 π R 2 m ( ξ , η ) 2 d ξ d η .

Moreover, if A 1 , A 2 , A n are HS operators with Fourier kernels m 1 , m 2 , m n , then

(2.3) tr ( A 1 A 2 A n ) = 1 2 π R n m 1 ( ξ 1 , ξ 2 ) m n ( ξ n , ξ 1 ) d ξ 1 d ξ n .

Lemma 2.1 also holds in the circle setting. The following results are elementary and will be used frequently later.

Lemma 2.2

(see Lemma 1.4 in [43]) Let A i denote HS operators on L 2 ( R ) with integral kernels K i ( x , y ) . Then,

(2.4) A i o p A i I 2 ,

(2.5) tr ( A 1 A ) i = 1 A i I 2 for all integers 2 .

Lemma 2.3

(See Lemma 1.5 in [43]) Let t A ( t ) define a C 1 curve in I 2 . Suppose

A ( t 0 ) I 2 < 1 3 ,

then there is a closed neighborhood I of t 0 on which the series

(2.6) α ( t ) = 2 ( 1 ) tr { A ( t ) }

converges and defines a C 1 function with

(2.7) d d t α ( t ) = 2 ( 1 ) tr A ( t ) 1 d d t A ( t ) .

Moreover, if A ( t ) is self-adjoint, then

(2.8) 1 3 A ( t ) I 2 2 α ( t ) 2 3 A ( t ) I 2 2 f o r a l l t I .

This lemma indicates that α ( t ) is differentiable and submits to term-by-term differentiation. If A ( t ) is self-adjoint, then α ( t ) is comparable to HS norm of A ( t ) .

3 Main estimates for the leading term

From Lemma 2.3 we know that the HS norm of operator A is almost conserved as long as the perturbation determinant is conserved under the assumption that A is self-adjoint (like the CH equation, A = R 0 q R 0 ). If A is not self-adjoint (like the FL equation), by studying the relation between the leading term in expansion of the perturbation determinant and the Sobolev norm of operator A , we can also obtain the increment of Sobolev norm of the solution.

In this section, we develop a new technique in Fourier analysis and complex analysis to calculate the leading term of the series expansion

(3.1) α FL ( κ ; q ( t ) ) = Re = 1 i 1 κ tr { [ ( κ ) 1 q x ( κ + ) 1 q ¯ x ] }

and

α CH ( κ ; q ( t ) ) = = 2 ( 1 ) tr { ( λ R 0 q R 0 ) } ,

even though they have been shown in [43,45,61] by using the explicit kernels of the free resolvent ( κ ± ) 1 and R 0 = ( 2 + κ 2 ) 1 . Our method is based on Fourier transform and the residue theorem in complex analysis. It is not limited to the explicit kernel of the operators, but also applicable to general operators, for instance, the higher order operators.

Proposition 3.1

(see Lemma 4.2 in [43] and Lemma 3.3 in [61]) Let κ > 0 . Assume that q is a Schwartz function on R , then we have

(3.2) Re tr { κ ( κ ) 1 q ( κ + ) 1 q ¯ } = R 2 κ 2 q ˆ ( ξ ) 2 4 κ 2 + ξ 2 d ξ .

If q is a Schwartz function on R Z , then

(3.3) Re tr { κ ( κ ) 1 q ( κ + ) 1 q ¯ } = 1 + e κ 1 e κ ξ 2 π Z 2 κ 2 q ˆ ( ξ ) 2 4 κ 2 + ξ 2 .

Proof

We provide different proofs for (3.2) and (3.3) by using Fourier transform method.

The line case. An elementary calculation gives

F ( ( κ ) 1 q ( κ + ) 1 q ¯ f ) ( ξ 3 ) = R 2 q ˆ ( ξ 3 ξ 2 ) κ i ξ 3 q ¯ ˆ ( ξ 2 ξ 1 ) κ + i ξ 2 f ˆ ( ξ 1 ) d ξ 1 d ξ 2 ,

then by Lemma 2.1 we have

tr { ( κ ) 1 q ( κ + ) 1 q ¯ } = 1 2 π R 2 q ˆ ( ξ 1 ξ 2 ) 2 ( κ i ξ 1 ) ( κ + i ξ 2 ) d ξ 1 d ξ 2 = 1 2 π R 2 ( κ 2 + ξ 1 ξ 2 ) q ˆ ( ξ 1 ξ 2 ) 2 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 2 2 ) d ξ 1 d ξ 2 + 1 2 π R 2 i κ ( ξ 1 ξ 2 ) q ˆ ( ξ 1 ξ 2 ) 2 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 2 2 ) d ξ 1 d ξ 2 .

Then,

(3.4) Re tr { ( κ ) 1 q ( κ + ) 1 q ¯ } = 1 2 π R 2 ( κ 2 + ξ 1 ξ 2 ) q ˆ ( ξ 1 ξ 2 ) 2 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 2 2 ) d ξ 1 d ξ 2 .

Taking variable substitution with ξ 1 + ξ 2 = 2 κ x and ξ 1 ξ 2 = 2 κ y , the right-hand side of (3.4) becomes

(3.5) 1 π R q ˆ ( 2 κ y ) 2 R 1 + x 2 y 2 ( 1 + ( x + y ) 2 ) ( 1 + ( x y ) 2 ) d x d y .

We use the residue theorem to calculate the inner integral in (3.5). As the integrand has no pole on the real axis, the integral is equal to 2 π i times the sum of the residues of the integrand in the upper half plane. If y = 0 , it is easy to see that the inner integral with respect to x in (3.5) is equal to π . If y R \ { 0 } , we denote

f y ( z ) = 1 + z 2 y 2 ( 1 + ( z + y ) 2 ) ( 1 + ( z y ) 2 ) , z C ,

which possesses two single poles z 1 = y + i and z 2 = y + i in the upper half plane. It is easy to see that

Res [ f y ( z ) ; z 1 ] = lim z z 1 ( z + y i ) f y ( z ) = 1 4 ( i y ) , Res [ f y ( z ) ; z 2 ] = lim z z 2 ( z y i ) f y ( z ) = 1 4 ( i + y ) ,

so we have

(3.6) R 1 + x 2 y 2 ( 1 + ( x + y ) 2 ) ( 1 + ( x y ) 2 ) d x = 2 π i k = 1 , 2 Res [ f y ( z ) ; z k ] = π y 2 + 1 .

Substituting (3.6) into (3.5), we obtain

Re tr { κ ( κ ) 1 q ( κ + ) 1 q ¯ } = κ R q ˆ ( 2 κ y ) 2 y 2 + 1 d y = R 2 κ 2 q ˆ ( ξ ) 2 4 κ 2 + ξ 2 d ξ .

The circle case. Similar to (3.4), we have

Re tr { ( κ ) 1 q ( κ + ) 1 q ¯ } = ξ 1 , ξ 2 2 π Z κ 2 + ξ 1 ξ 2 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 2 2 ) q ˆ ( ξ 1 ξ 2 ) 2 .

Taking η = ξ 1 ξ 2 gives that

(3.7) Re tr { ( κ ) 1 q ( κ + ) 1 q ¯ } = η 2 π Z ξ 1 2 π Z κ 2 + ξ 1 ( ξ 1 η ) ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 1 η 2 ) q ˆ ( η ) 2 .

Denote

f ( ξ 1 ) = κ 2 + ξ 1 ( ξ 1 η ) ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 1 η 2 ) , ξ 1 , η 2 π Z ,

we can use the Poisson summation formula

(3.8) ξ 1 2 π Z f ( ξ 1 ) = 1 2 π ζ Z R f ( x ) e i x ζ d x

to calculate the inner summation in the right-hand side of (3.7).

If η = 0 , from (3.8) we need to calculate the integral

I ( ζ ) = R 1 κ 2 + x 2 e i ζ x d x , ζ Z .

From the Fourier transform of the well-known Poisson kernel, we can of course know that I ( ζ ) = π κ e κ ζ . However, in order to introduce our method more clearly, we recalculate it via the residue theorem. Write

H ( z ) = e i ζ z κ 2 + z 2 , z C ,

then H ( z ) is analytic near the real line. From Jordon’s lemma, we know that

lim R + Γ R e i ζ z κ 2 + z 2 d z = 0 ( for ζ 0 ) , and lim R + Γ R e i ζ z κ 2 + z 2 d z = 0 ( for ζ > 0 ) ,

where Γ R { z C z = Re i θ , θ [ 0 , π ] } , Γ R { z C z = Re i θ , θ [ π , 0 ] } . Then, I ( ζ ) reduces to obtain the residues in the upper half plane or the lower half plane. Note that z = κ i and z = κ i are the single poles of H ( z ) , and we have

Res [ H ( z ) ; κ i ] = e κ ζ 2 κ i , Res [ H ( z ) ; κ i ] = e κ ζ 2 κ i .

Hence, we can obtain from the residue theorem that

(3.9) I ( ζ ) = R e i ζ x κ 2 + x 2 d x = 2 π i Res [ H ( z ) ; κ i ] = π κ e κ ζ , ( for ζ 0 ) , I ( ζ ) = R e i ζ x κ 2 + x 2 d x = 2 π i Res [ H ( z ) ; κ i ] = π κ e κ ζ , ( for ζ > 0 ) .

Therefore, we obtain the conclusion from (3.8) that

(3.10) ξ 1 2 π Z 1 κ 2 + ξ 1 2 = 1 2 π ζ Z π κ e κ ζ = 1 + e κ 2 κ ( 1 e κ ) .

If η 2 π Z \ { 0 } , we need to calculate the integral

I ( ζ ) = R κ 2 + x ( x η ) ( κ 2 + x 2 ) ( κ 2 + x η 2 ) e i ζ x d x .

Similar to the previous case, we write

H η ( z ) = κ 2 + z ( z η ) ( κ 2 + z 2 ) ( κ 2 + z η 2 ) e i ζ z , z C , ζ Z , η 2 π Z \ { 0 } ,

and know that the poles of H ( z ) located in the upper half plane and the lower half plane are z 1 = κ i , z 2 = η + κ i , and z 3 = κ i , z 4 = η κ i , respectively. A simple calculation gives

Res [ H η ( z ) ; z 1 ] = e κ ζ 2 ( 2 κ i η ) , Res [ H η ( z ) ; z 2 ] = e ζ ( κ i η ) 2 ( 2 κ i + η ) = e κ ζ 2 ( 2 κ i + η ) , Res [ H η ( z ) ; z 3 ] = e κ ζ 2 ( 2 κ i + η ) , Res [ H η ( z ) ; z 4 ] = e ζ ( κ + i η ) 2 ( 2 κ i η ) = e κ ζ 2 ( 2 κ i η ) ,

which imply that

(3.11) I ( ζ ) = 2 π i k = 1 , 2 Res [ H η ( z ) ; z k ] = 4 κ π 4 κ 2 + η 2 e κ ζ , ( for ζ 0 ) ,

(3.12) I ( ζ ) = 2 π i k = 3 , 4 Res [ H η ( z ) ; z k ] = 4 κ π 4 κ 2 + η 2 e κ ζ , ( for ζ > 0 ) .

Plugging (3.11) and (3.12) into the right-hand side of (3.8), one can obtain that

(3.13) ξ 1 2 π Z κ 2 + ξ 1 ( ξ 1 η ) ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 1 η 2 ) = 2 κ 4 κ 2 + η 2 ζ Z e κ ζ = 2 κ 4 κ 2 + η 2 1 + e κ 1 e κ .

Finally, substituting (3.10) and (3.13) into the right-hand side of (3.7), we obtain

Re tr { κ ( κ ) 1 q ( κ + ) 1 q ¯ } = 1 + e κ 1 e κ ξ 2 π Z 2 κ 2 q ˆ ( ξ ) 2 4 κ 2 + ξ 2 .

The proof is completed.□

Proposition 3.2

(Proposition 2.1 in [43]) Let κ > 0 , R 0 = ( x 2 + κ 2 ) 1 . Assume that q is a real-valued Schwartz function, then we have

(3.14) R 0 q R 0 I 2 ( R ) 2 = 1 κ R q ˆ ( ξ ) 2 ξ 2 + 4 κ 2 d ξ ,

(3.15) R 0 q R 0 I 2 ( R Z ) 2 = 2 e κ ( 1 e κ ) 2 q ˆ ( 0 ) 2 4 κ 2 + 1 e 2 κ κ ( 1 e κ ) 2 ξ 2 π Z q ˆ ( ξ ) 2 ξ 2 + 4 κ 2 .

Proof

Here we give an alternative proof.

The line case. Since

F ( ( R 0 q ) 2 f ) ( ξ 3 ) = R 2 q ˆ ( ξ 3 ξ 2 ) κ 2 + ξ 3 2 q ˆ ( ξ 2 ξ 1 ) κ 2 + ξ 2 2 f ˆ ( ξ 1 ) d ξ 1 d ξ 2 ,

and R 0 q R 0 is self-adjoint, then it follows from Lemma 2.1 that

(3.16) R 0 q R 0 I 2 ( R ) 2 = tr ( R 0 q R 0 R 0 q R 0 ) = tr ( ( R 0 q ) 2 ) = 1 2 π R 2 q ˆ ( ξ 1 ξ 2 ) 2 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 2 2 ) d ξ 1 d ξ 2 .

Taking ξ 1 + ξ 2 = 2 κ x and ξ 1 ξ 2 = 2 κ y , the right-hand side of (3.16) can be written as

(3.17) 1 π κ 2 R q ˆ ( 2 κ y ) 2 R 1 ( 1 + ( x + y ) 2 ) ( 1 + ( x y ) 2 ) d x d y .

Similar to the above proposition, the residue theorem is used to calculate the internal integral in (3.17). If y = 0 , it is easy to see that

R 1 ( 1 + x 2 ) 2 d x = 2 π i Res 1 ( 1 + z 2 ) 2 ; i = π 2 .

If y R \ { 0 } , we denote

f y ( z ) = 1 ( 1 + ( z + y ) 2 ) ( 1 + ( z y ) 2 ) , z C .

f y ( z ) possesses two single poles z 1 = y + i and z 2 = y + i in the upper half plane. Note that

Res [ f y ( z ) ; z 1 ] = 1 8 i y ( y i ) , Res [ f y ( z ) ; z 2 ] = 1 8 i y ( y + i ) ,

hence

(3.18) R 1 ( 1 + ( x + y ) 2 ) ( 1 + ( x y ) 2 ) d x = 2 π i k = 1 , 2 Res [ f y ( z ) ; z k ] = π 2 ( y 2 + 1 ) .

Substituting (3.18) into (3.17), we can obtain the conclusion (3.14).

The circle case. The HS norm of R 0 q R 0 in the circle setting can be expressed as

R 0 q R 0 I 2 ( R Z ) 2 = ξ 1 , ξ 2 2 π Z q ˆ ( ξ 1 ξ 2 ) 2 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 2 2 ) .

Setting η = ξ 1 ξ 2 , then

(3.19) R 0 q R 0 I 2 ( R Z ) 2 = η 2 π Z ξ 1 2 π Z 1 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 1 η 2 ) q ˆ ( η ) 2 .

We use the Poisson summation formula (3.8) to calculate the inner summation in right-hand side of (3.19). Denote

f η ( ξ 1 ) = 1 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 1 η 2 ) , I ( ζ ) = 1 2 π R 1 ( κ 2 + x 2 ) ( κ 2 + x η 2 ) e i ζ x d x .

Then, by using the residue theorem, we obtain

I ( ζ ) = 1 2 π 2 π i Res e i ζ z ( κ 2 + z 2 ) 2 ; κ i = 1 4 κ 3 + ζ 4 κ 2 e κ ζ , if η = 0 .

Similar to (3.11) and (3.12), we know that

I ( ζ ) = 1 κ ( η 2 + 4 κ 2 ) e κ ζ , if η 0 .

Plugging these identities into (3.8), one can obtain that

(3.20) ξ 1 2 π Z 1 ( κ 2 + ξ 1 2 ) 2 = ( 1 + e κ ) 4 κ 3 ( 1 e κ ) + 2 e κ 4 κ 2 ( 1 e κ ) 2

and

(3.21) ξ 1 2 π Z 1 ( κ 2 + ξ 1 2 ) ( κ 2 + ξ 1 η 2 ) = 1 + e κ κ ( η 2 + 4 κ 2 ) ( 1 e κ ) if η 0 .

Substituting (3.20) and (3.21) into the right-hand side of (3.19), we obtain the conclusion (3.15). The proof is now complete.□

Corollary 3.3

If u H 1 is a real-valued function on R or R Z , λ > 0 and κ 2 = 1 4 + λ . Denote q = u u x x , then

(3.22) 1 κ q , ( x 2 + 4 κ 2 ) 1 q R 0 q R 0 I 2 2 10 κ q , ( x 2 + 4 κ 2 ) 1 q .

Consequently, if λ 1 , then we have

(3.23) 3 2 u H 1 2 R 0 q R 0 I 2 2 21 u H 1 2 .

Proof

(3.22) comes immediately from (3.14) and (3.15). If λ 1 , κ 2 = 1 4 + , thus (3.23) is a direct consequence of (3.3).□

4 Conservation of the perturbation determinant

To guarantee convergence and term-by-term differentiation of the series appearing in α FL ( κ ; q ( t ) ) , we need the following estimates.

Lemma 4.1

(See Lemma 3.2 in [45]) Let q S , l 2 , and κ > 0 . Then,

tr { [ κ ( κ ) 1 q ( κ + ) 1 q ¯ ] } ( κ s q H s ) 2 l ( 0 s < 1 4 ) , ( κ 1 4 q H s ) 2 l ( s 1 4 ) .

The next two theorems show that the perturbation determinant is conserved along the flows generated by FL and CH, respectively.

Theorem 4.2

Let q ( t ) be a Schwartz solution to (1.7). Assume that q x ( 0 ) L 2 c 1 or κ q x H s 1 a for s > 0 where a = min ( 1 4 , s ) . Then,

d d t α FL ( κ ; q ( t ) ) = 0 .

Proof

Let us write R = ( κ ) 1 and R + = ( + κ ) 1 for simplicity. As q ( t ) is a Schwartz solution to the FL equation, it suffices to show that

= 1 ( i κ ) tr { ( R q x R + q ¯ x ) 1 [ R q t x R + q ¯ x + R q x R + q ¯ t x ] } = = 1 ( A B C ) = 0 ,

where

A = ( i κ ) tr { ( R q x R + q ¯ x ) 1 [ R q x x R + q ¯ x + R q x R + q ¯ x x ] } , B = ( i κ ) tr { ( R q x R + q ¯ x ) 1 [ R ( i q 2 q x ) R + q ¯ x R q x R + ( i q 2 q ¯ x ) ] } , C = ( i κ ) tr { ( R q x R + q ¯ x ) 1 [ R q R + q ¯ x + R q x R + q ¯ ] } .

Note that

q x x = R 1 q x q x R + 1 + 2 κ q x , q ¯ x x = R + 1 q ¯ x q ¯ x R 1 2 κ q ¯ x .

Putting these identities into A , we obtain

(4.1) A = ( i κ ) tr { ( R q x R + q ¯ x ) 1 [ q x R + q ¯ x R q x 2 + 2 κ R q x R + q ¯ x + R q x 2 R q x R + q ¯ x ( κ ) 2 κ R q x R + q ¯ x ] } .

By cycling the trace,

tr { ( R q x R + q ¯ x ) 1 R q x R + q ¯ x ( κ ) } = tr { ( κ ) ( R q x R + q ¯ x ) } = tr { q x R + q ¯ x ( R q x R + q ¯ x ) 1 } = tr { ( R q x R + q ¯ x ) 1 q x R + q ¯ x } ,

which thus cancels the first term in the right-hand side of (4.1). Sequentially, A = 0 for any 1 .

Hence, it suffices to prove that C 1 = 0 and

B + C + 1 = 0 for 1 ,

which is also equivalent to

(4.2) tr { ( R q x R + q ¯ x ) 1 ( R q 2 q x R + q ¯ x R q x R + q 2 q ¯ x ) } = κ tr { ( R q x R + q ¯ x ) ( R q R + q ¯ x + R q x R + q ¯ ) }

for 1 .

On the other hand, observe that

q 2 R 1 R 1 q 2 = [ q 2 , κ ] = ( q 2 ) x ,

pre-multiplying and post-multiplying, respectively, each term above by R gives

(4.3) R q 2 q 2 R = R ( q 2 ) x R .

Hence, by using (4.3), one obtains that

(4.4) LHS ( 4.2 ) = tr { ( R q x R + q ¯ x ) 1 ( R q 2 q x R + q ¯ x q 2 R q x R + q ¯ x ) } + tr { ( R q x R + q ¯ x ) 1 ( q 2 R q x R + q ¯ x R q x R + q 2 q ¯ x ) } = tr { ( R q x R + q ¯ x ) R ( q 2 ) x } .

The second term in the penultimate line above vanishes after additionally cycling the trace. Using the elementary operator identities:

(4.5) q x = R 1 q q R + 1 + 2 κ q , q ¯ x = R + 1 q ¯ q ¯ R 1 2 κ q ¯ ,

we deduce that

(4.6) R q x R + q ¯ x R q ¯ x q + R q ¯ q x R q x R + q ¯ x = R q x R + ( R + 1 q ¯ q ¯ R 1 2 κ q ¯ ) R q ¯ x q + R q ¯ q x R ( R 1 q q R + 1 + 2 κ q ) R + q ¯ x = R q x q ¯ R q ¯ x q R q x R + q ¯ q ¯ x q 2 κ R q x R + q ¯ R q ¯ x q + R q ¯ q x q R + q ¯ x R q ¯ q x R q q ¯ x + 2 κ R q ¯ q x R q R + q ¯ x = 2 κ R q ¯ q x R q R + q ¯ x 2 κ R q x R + q ¯ R q ¯ x q + R q 2 q x R + q ¯ x R q x R + q 2 q ¯ x .

Combining (4.6) with (4.3) yields

tr { ( R q x R + q ¯ x ) R ( q 2 ) x } = tr { ( R q x R + q ¯ x ) 1 ( R q x R + q ¯ x R q ¯ x q + R q ¯ q x R q x R + q ¯ x ) } = 2 κ tr { ( R q x R + q ¯ x ) 1 ( R q ¯ q x R q R + q ¯ x R q x R + q ¯ R q ¯ x q ) } + tr { ( R q x R + q ¯ x ) 1 ( R q 2 q x R + q ¯ x R q x R + q 2 q ¯ x ) } = 2 κ tr { ( R q x R + q ¯ x ) 1 ( R q ¯ q x R q R + q ¯ x R q x R + q ¯ R q ¯ x q ) } + tr { ( R q x R + q ¯ x ) 1 ( q 2 R q x R + q ¯ x R q x R + q 2 q ¯ x ) } tr { ( R q x R + q ¯ x ) 1 R ( q 2 ) x R q x R + q ¯ x } = 2 κ tr { ( R q x R + q ¯ x ) 1 ( R q ¯ q x R q R + q ¯ x R q x R + q ¯ R q ¯ x q ) } tr { ( R q x R + q ¯ x ) R ( q 2 ) x } ,

which further gives

tr { ( R q x R + q ¯ x ) R ( q 2 ) x } = κ tr { ( R q x R + q ¯ x ) 1 ( R q ¯ q x R q R + q ¯ x R q x R + q ¯ R q ¯ x q ) } .

Plugging this identity into (4.4), we obtain

(4.7) LHS ( 4.2 ) = κ tr { ( R q x R + q ¯ x ) 1 ( R q x R + q ¯ R q ¯ x q R q ¯ q x R q R + q ¯ x ) } .

On the other hand, we find that

(4.8) RHS ( 4.2 ) = κ tr { ( R q x R + q ¯ x ) 1 ( R q x R + q ¯ x R q R + q ¯ x + R q x R + q ¯ R q x R + q ¯ x ) } = κ tr { ( R q x R + q ¯ x ) 1 R q x R + ( R + 1 q ¯ q ¯ R 1 2 κ q ¯ ) R q R + q ¯ x } κ tr { ( R q x R + q ¯ x ) 1 R q x R + q ¯ R ( R 1 q q R + 1 + 2 κ q ) R + q ¯ x } = κ tr { ( R q x R + q ¯ x ) 1 ( R q x q ¯ R q R + q ¯ x R q x R + q 2 R + q ¯ x 2 κ R q x R + q ¯ R q R + q ¯ x ) } κ tr { ( R q x R + q ¯ x ) 1 ( R q x R + q 2 R + q ¯ x R q x R + q ¯ R q q ¯ x + 2 κ R q x R + q ¯ R q R + q ¯ x ) } = κ tr { ( R q x R + q ¯ x ) 1 ( R q x R + q ¯ R q ¯ x q R q ¯ q x R q R + q ¯ x ) } .

Thus, the veracity of (4.2) follows readily from (4.7) and (4.8).

In the end, we compute C 1 from Fourier transform side (see Lemma 2.1)

C 1 = i κ tr { R q R + q ¯ x + R q x R + q ¯ } = κ ( ξ 2 ξ 1 ) q ˆ ( ξ 1 ξ 2 ) 2 ( i ξ 1 κ ) ( i ξ 2 + κ ) d ξ 1 d ξ 2 + ( ξ 1 ξ 2 ) q ˆ ( ξ 1 ξ 2 ) 2 ( i ξ 1 κ ) ( i ξ 2 + κ ) d ξ 1 d ξ 2 = 0 .

The proof is complete.□

Remark 4.3

We can also utilize the tool of operator trace to compute C 1 . In fact, using (4.5) and cycling the trace, one has

C 1 = i κ tr { R q R + q ¯ x + R q x R + q ¯ } = i κ tr { R q R + ( R + 1 q ¯ q ¯ R 1 2 κ q ¯ ) + R ( R 1 q q R + 1 + 2 κ q ) R + q ¯ } = i κ tr { R q 2 R q R + q ¯ R 1 2 κ R q R + q ¯ + q R + q ¯ R q 2 + 2 κ R q R + q ¯ } = i κ tr { q R + q ¯ R q R + q ¯ R 1 } = 0 .

Theorem 4.4

Let u ( t ) be a Schwartz solution to (1.2) with ω = 1 . Denote κ 2 = 1 4 + λ , and q = u u x x . Then,

d d t α CH ( κ ; q ( t ) ) = 0

for λ u H 1 1 .

Proof

From (3.23), we have

λ R 0 q R 0 I 2 2 21 λ 2 u H 1 2 1 ,

which shows that Lemma 2.3 applies. u ( t ) is a Schwartz solution to CH, thus

d d t α CH ( κ ; q ( t ) ) = = 2 ( 1 ) λ tr { ( R 0 q R 0 ) 1 R 0 q t R 0 } = = 2 ( 1 ) λ tr { ( R 0 q ) 1 R 0 q t } = = 2 ( 1 ) λ tr { ( R 0 q ) 1 R 0 ( 2 u x + 3 u u x 2 u x u x x u u x x x ) } .

In view of the above, it suffices to show

(4.9) 2 λ tr { ( R 0 q ) R 0 u x } = tr { ( R 0 q ) 1 R 0 ( 3 u u x 2 u x u x x u u x x x ) } for 2 ,

(4.10) and tr { R 0 q R 0 u x } = 0 .

By commuting R 0 and and then cycling the trace, one obtains

tr { ( R 0 q ) R 0 [ , q ] } = tr { ( R 0 q ) R 0 q } tr { ( R 0 q ) R 0 q } = tr { ( R 0 q ) R 0 q } tr { ( R 0 q ) + 1 } = 0 ,

which implies that

(4.11) tr { ( R 0 q ) R 0 u x } = tr { ( R 0 q ) R 0 u x x x } .

Substituting the operator identity

u x x x = R 0 1 u x u x R 0 1 + 2 R 0 1 u 2 u R 0 1 + 4 κ 2 [ , u ]

into (4.11), we have

tr { ( R 0 q ) R 0 u x } = tr { ( R 0 q ) u x } tr { ( R 0 q ) R 0 u x R 0 1 } + 2 tr { ( R 0 q ) u } 2 tr { ( R 0 q ) R 0 u R 0 1 } + 4 κ 2 tr { ( R 0 q ) R 0 u x } = tr { ( R 0 q ) u x } tr { ( R 0 q ) 1 R 0 u x R 0 1 R 0 q } + 2 tr { ( R 0 q ) u } 2 tr { ( R 0 q ) 1 R 0 u R 0 1 R 0 q } + 4 κ 2 tr { ( R 0 q ) R 0 u x } = 2 tr { ( R 0 q ) u x } 2 tr { ( R 0 q ) 1 R 0 [ , u q ] } + 4 κ 2 tr { ( R 0 q ) R 0 u x } .

Hence,

2 tr { ( R 0 q ) u x } + 2 tr { ( R 0 q ) 1 R 0 [ , u q ] } = ( 4 κ 2 1 ) tr { ( R 0 q ) R 0 u x } = 4 λ tr { ( R 0 q ) R 0 u x } .

Therefore, we can obtain that

2 λ tr { ( R 0 q ) R 0 u x } = tr { ( R 0 q ) 1 R 0 q u x } + tr { ( R 0 q ) 1 R 0 [ , u q ] } = tr { ( R 0 q ) 1 R 0 ( 2 q u x + u q x ) } = tr { ( R 0 q ) 1 R 0 [ 2 ( u u x x ) u x + u ( u x u x x x ) ] } = tr { ( R 0 q ) 1 R 0 ( 3 u u x 2 u x u x x u u x x x ) } ,

which yields (4.9).

The identity (4.10) is easily verified. Note that

tr { R 0 f R 0 f x } = tr { R 0 f R 0 [ , f ] } = tr { R 0 f R 0 f } tr { ( R 0 f ) 2 } = 0 .

By taking f = u and f = u x , respectively, one obtains

tr { R 0 q R 0 u x } = tr { R 0 u R 0 u x } tr { R 0 u x x R 0 u x } = 0 .

The proof of this theorem is now complete.□

5 Conservation of norms for FL and CH

In this section, our first task is to obtain bounds on Besov norms of solution to the FL equation in the range 0 < s < 1 2 . The key is to construct Besov norms of solution to the FL equation from the leading term of α FL ( κ ; q ( t ) ) (see (3.2) and (3.3)). Even though, Proposition 3.1 indicates that the leading term can be treated as q x H 1 , this norm is not equivalent to q L 2 . Moreover, the function

ξ κ 2 4 κ 2 + ξ 2

does not decay as ξ 0 . In fact, we would need a function decaying suitably as ξ 0 when the regularity index s is non-negative. Hence, we take linear combinations of these functions at different values of κ to obtain more decay.

Lemma 5.1

(See Lemma 3.5 in [43]) Fix 1 r , 1 < s < 1 and define

w ( ξ , κ ) = κ 2 ξ 2 + 4 κ 2 ( κ 2 ) 2 ξ 2 + κ 2 = 3 κ 2 ξ 2 4 ( ξ 2 + κ 2 ) ( ξ 2 + 4 κ 2 ) .

Then,

(5.1) f B 2 , r s s f H 1 + κ 0 N 2 N N r s f , w ( i x , κ 0 N ) f r 2 1 r

and

(5.2) N 2 N N r s f , w ( i x , κ 0 N ) f r 2 1 r κ 0 s f B 2 , r s

uniformly for κ 0 1 .

Remark 5.2

f H 1 in the right-hand side of (5.1) is used to control the low frequency part of f B 2 , r s (i.e., f ˆ ( ξ ) L 2 ( ξ 1 ) ). For the KdV equation, if q ( t ) is a smooth solution, q ( t ) H 1 was proved to be almost conserved. Hence, Lemma 5.1 tells us that conservation of q ( t ) B 2 , r s comes from conservation of q , w ( i x , κ 0 N ) q . Further, the term q , κ 2 x 2 + 4 κ 2 q is closely related to the leading term of perturbation determinant. Therefore, one can obtain conservation of Besov norm of the solution from conservation of perturbation determinant. Since the regularity we consider here is non-negative, and according to the known conservation laws (see (1.8)) of the FL equation, we will use the following estimate:

(5.3) q x ( t ) B 2 , r s s q x ( 0 ) L 2 + κ 0 N 2 N N r s q x ( t ) , w ( i x , κ 0 N ) q x ( t ) r 2 1 r .

Then, we give the proof of the main theorem about the almost conservation laws for the FL equation.

Proof of Theorem 1.1

We begin with the line case. In view of (3.2), we obtain

q x , κ 2 x 2 + 4 κ 2 q x = 1 2 Re ( tr { κ ( κ ) 1 q x ( κ + ) 1 q ¯ x } ) 1 2 α 1 ( κ ; q ( t ) ) .

Note that

q x ( t ) , w ( i x , κ ) q x ( t ) = q x , κ 2 x 2 + 4 κ 2 q x q x , ( κ 2 ) 2 x 2 + κ 2 q x ,

Then, we obtain that

(5.4) q x ( t ) , w ( i x , κ ) q x ( t ) = 1 2 α 1 ( κ ; q ( t ) ) 1 2 α 1 ( κ 2 ; q ( t ) ) ,

which implies that

(5.5) q x ( t ) , w ( i x , κ ) q x ( t ) q x ( 0 ) , w ( i x , κ ) q x ( 0 ) α 1 ( κ ; q ( t ) ) α 1 ( κ ; q ( 0 ) ) + α 1 ( κ 2 ; q ( t ) ) α 1 ( κ 2 ; q ( 0 ) ) .

According to Lemma 4.1, we have

(5.6) α ( κ ; q ( t ) ) α 1 ( κ ; q ( t ) ) = 2 tr { [ κ R q x R + q ¯ x ] } κ 4 s q x ( t ) H s 4 ,

provided that 0 < s < 1 4 and κ s q x ( t ) H s 1 . Combining (5.5), (5.6), and Theorem 4.2, we see that

(5.7) q x ( t ) , w ( i x , κ ) q x ( t ) q x ( 0 ) , w ( i x , κ ) q x ( 0 ) + κ 4 s ( q x ( t ) H s 4 + q x ( 0 ) H s 4 ) q x ( 0 ) , w ( i x , κ ) q x ( 0 ) + κ 4 s ( q x ( t ) B 2 , r s 4 + q x ( 0 ) B 2 , r s 4 ) .

The veracity of the last inequality follows readily from the embedding B 2 , r s H s for s > s and r 1 . Recalling (5.3), we denote

(5.8) f Z κ 0 N 2 N N r s f , w ( i x , κ 0 N ) f r 2 1 r ,

and D s the constant such that

(5.9) q x ( t ) B 2 , r s D s ( q x ( 0 ) L 2 + κ 0 q x ( t ) Z κ 0 ) .

Then, from (5.7) with κ = κ 0 N , we have

(5.10) q x ( t ) Z κ 0 q x ( 0 ) Z κ 0 + N 2 N N r s ( κ 0 N ) 2 s r 1 r ( q x ( t ) B 2 , r s 2 + q x ( 0 ) B 2 , r s 2 ) q x ( 0 ) Z κ 0 + κ 0 2 s ( q x ( t ) B 2 , r s 2 + q x ( 0 ) B 2 , r s 2 )

by choosing s < s < 2 s . This is practicable because 0 < s < 1 2 .

Taking

κ 0 s = 100 ( 1 + q x ( 0 ) B 2 , r s )

such that

κ 0 s q x ( 0 ) H s κ 0 s q x ( 0 ) B 2 , r s 1 .

Correspondingly, by using (5.9) and (5.10), we obtain

(5.11) q x ( t ) Z κ 0 q x ( 0 ) Z κ 0 + κ 0 2 s ( κ 0 2 q x ( t ) Z κ 0 2 + κ 0 2 q x ( 0 ) Z κ 0 2 + q x ( 0 ) L 2 2 )

for t I , a small neighborhood of the temporal origin.

In order to use the bootstrap argument, we rewrite (5.11) as

(5.12) κ 0 2 ( 1 s ) q x ( t ) Z κ 0 C s ( κ 0 2 ( 1 s ) q x ( 0 ) Z κ 0 + κ 0 2 ( 1 2 s ) q x ( 0 ) L 2 2 + ( κ 0 2 ( 1 s ) q x ( t ) Z κ 0 ) 2 + ( κ 0 2 ( 1 s ) q x ( 0 ) Z κ 0 ) 2 )

with C s > 1 . Suppose that

(5.13) max ( κ 0 2 ( 1 s ) q x ( 0 ) Z κ 0 , κ 0 ( 1 2 s ) q x ( 0 ) L 2 ) ε 1 ,

where

κ 0 = [ 100 ( 1 + q x ( 0 ) B 2 , r s ) ] 1 s .

We will show that

(5.14) κ 0 2 ( 1 s ) q x ( t ) Z κ 0 2 C s ε

for any t R . To this end, let I be the maximal interval containing the origin such that (5.14) holds true for any t I . Due to the continuity of q x ( t ) Z κ 0 , we know that I is non-empty. Moreover, I is open. (5.12), (5.14), and the assumption on q x ( 0 ) Z κ 0 and q x ( 0 ) L 2 yield

(5.15) κ 0 2 ( 1 s ) q x ( t ) Z κ 0 C s ( ε + 2 ε 2 + ( 2 C s ε ) 2 ) 3 2 C s ε

by choosing ε < ( 16 C s 2 ) 1 . To guarantee the convergence of series, we additionally choose ε small enough such that

q x ( t ) H s q x ( t ) B 2 , r s D s ε + 3 2 C s ε < 1 2

by (5.9) and (5.15). Then, continuity argument yields I = R .

(5.14) tells us that

q x ( t ) Z κ 0 2 C s ( q x ( 0 ) Z κ 0 + κ 0 1 q x ( 0 ) L 2 ) ,

which further implies from (5.2) and (5.3) that

q x ( t ) B 2 , r s q x ( 0 ) L 2 + κ 0 1 s q x ( 0 ) B 2 , r s q x ( 0 ) B 2 , r s .

By an analysis of the power of κ 0 in (5.13), which is positive, we cannot obtain smallness by choosing κ 0 sufficiently large. Besides, there is no scaling symmetry for the FL equation, so the initial data with small Besov norm is essential. This finishes the proof of Theorem 1.1 in the line case.

Now, we turn to the circle case. From (3.3), we know that

q x , κ 2 x 2 + 4 κ 2 q x = 1 e κ 2 ( 1 + e κ ) Re ( tr { κ ( κ ) 1 q x ( κ + ) 1 q ¯ x } ) 1 e κ 2 ( 1 + e κ ) α 1 ( κ ; q ( t ) ) .

Then,

q x ( t ) , w ( i x , κ ) q x ( t ) = 1 e κ 2 ( 1 + e κ ) α 1 ( κ ; q ( t ) ) 1 e κ 2 2 ( 1 + e κ 2 ) α 1 ( κ 2 ; q ( t ) ) .

Proceeding in the manner as the line case, we can obtain from Lemma 4.1 and Theorem 4.2 that

q x ( t ) , w ( i x , κ ) q x ( t ) q x ( 0 ) , w ( i x , κ ) q x ( 0 ) + κ 4 s ( q x ( t ) B 2 , r s 4 + q x ( 0 ) B 2 , r s 4 ) .

Furthermore, from (5.3), we have

(5.16) q x ( t ) Z κ 0 q x ( 0 ) Z κ 0 + κ 0 2 s ( q x ( t ) B 2 , r s 2 + q x ( 0 ) B 2 , r s 2 ) q x ( 0 ) Z κ 0 + κ 0 2 s ( κ 0 2 q x ( t ) Z κ 0 2 + κ 0 2 q x ( 0 ) Z κ 0 2 + q x ( 0 ) L 2 2 ) .

Then, a standard bootstrap argument implies that

q x ( t ) Z κ 0 q x ( 0 ) Z κ 0

uniformly in time. Applying Lemma 5.1 once again, we obtain

q x ( t ) B 2 , r s q x ( 0 ) L 2 + κ 0 1 s q x ( 0 ) B 2 , r s q x ( 0 ) B 2 , r s

for all t R . The proof of Theorem 1.1 is complete.□

In the end, we obtain in passing the result for the CH equation, that is, conservation of the series (1.15) under the CH flow guarantees global control on the H 1 -norm of the solution.

Proof of Theorem 1.4

In view of (3.23), by taking λ u ( 0 ) H 1 1 , we can see that

λ R 0 q ( 0 ) R 0 I 2 2 21 λ 2 u ( 0 ) H 1 2 1 .

So Lemma 2.3 applies. Combining with Theorem 4.4, there exists a closed neighborhood I of t = 0 on which

d d t α C H ( κ ; q ( t ) ) = 0 .

Therefore, by conservation of α C H ( κ ; q ( t ) ) and (2.8), we have

(5.17) λ 2 R 0 q ( t ) R 0 I 2 2 3 α C H ( κ ; q ( t ) ) = 3 α C H ( κ ; q ( 0 ) ) 2 λ 2 R 0 q ( 0 ) R 0 I 2 2 1

for t I . Therefore, standard continuity argument implies that (5.17) holds true for all t R . Applying (3.23) and (5.17), we obtain

(5.18) u ( t ) H 1 < R 0 q ( t ) R 0 I 2 2 R 0 q ( 0 ) R 0 I 2 < 10 u ( 0 ) H 1 .

The lower bound in (1.16) follows directly from the above upper bound by employing the time translation symmetry, which completes the proof of the theorem.□

Acknowledgements

The authors would like to thank Professor Zhong Wang (Foshan University) for useful discussion on conservation laws for the Camassa-Holm equation. They would especially like to thank the anonymous reviewers for their suggestions that have improved the presentation of this study.

  1. Funding information: M. Chen is supported by National Natural Science Foundation of China 12001236 and Natural Science Foundation of Guangdong Province 2020A1515110494. M. Shan is partially supported by NSFC grant 12101629.

  2. Author contributions: M. Shan and M. Chen developed the theoretical analyses. M. Shan performed the analytic calculations and drafted the manuscript. Y. Lu and J. Wang respectively calculated the conservation of the perturbation determinant for Fokas-Lenells equation and Camassa-Holm equation. M. Shan and M. Chen contributed to the final version of the article.

  3. Conflict of interest: The authors state no conflict of interest.

References

[1] M. S. Alber, R. Camassa, D. D. Holm, and J. E. Marsden, The geometry of peaked solitons and billiard solutions of a class of integrable PDEs, Lett. Math. Phys. 32 (1994), 137–151. 10.1007/BF00739423Search in Google Scholar

[2] H. Bahouri and G. Perelman, Global well-posedness for the derivative nonlinear Schrödinger equation, Invent. Math. 229 (2022), no. 2, 639–688. 10.1007/s00222-022-01113-0Search in Google Scholar

[3] H. A. Biagioni and F. Linares, Ill-posedness for the derivative Schrödinger and generalized Benjamin-Ono equations, Trans. Amer. Math. Soc. 353 (2001), no. 9, 3649–3659. 10.1090/S0002-9947-01-02754-4Search in Google Scholar

[4] A. Bressan and A. Constantin, Global solutions of the Hunter-Saxton equation, SIAM J. Math. Anal. 37 (2005), 996–1026. 10.1137/050623036Search in Google Scholar

[5] A. Bressan, G. Chen, and Q. Zhang, Uniqueness of conservative solutions to the Camassa-Holm equation via characteristics, Discrete Contin. Dyn. Syst. 35 (2015), no. 1, 25–42. 10.3934/dcds.2015.35.25Search in Google Scholar

[6] F. Calogero and J. P. Francoise, A completely integrable Hamiltonian system, J. Math. Phys. 37 (1996), 2863–2871. 10.1063/1.531536Search in Google Scholar

[7] R. Camassa and D. Holm, An integrable shallow water equation with peaked solitons, Phys. Rev. Lett. 71 (1993), 1661–1664. 10.1103/PhysRevLett.71.1661Search in Google Scholar PubMed

[8] R. Camassa, D. Holm, and J. Hyman, A new integrable shallow water equation, Adv. Appl. Math. 31 (1994), 1–33. 10.1016/S0065-2156(08)70254-0Search in Google Scholar

[9] M. Chen and M. Shan, A priori estimates for the fifth-order modified KdV equations in Besov spaces with low regularity, J. Appl. Anal. Comput. 13 (2023), no. 4, 2315–2325. 10.11948/20220538Search in Google Scholar

[10] R. Chen, W. Lian, D. Wang, and R. Xu, A rigidity property for the Novikov equation and the asymptotic stability of peakons, Archive Rational Mech. Anal. 241 (2021), 497–533. 10.1007/s00205-021-01658-zSearch in Google Scholar

[11] Q. Cheng and E. Fan, A Riemann-Hilbert approach to existence of global solutions to the Fokas-Lenells equation on the line, 2022, arXiv:2206.02155. 10.2139/ssrn.4677933Search in Google Scholar

[12] Q. Cheng and E. Fan, Long-time asymptotics for the focusing Fokas-Lenells equation in the solitonic region of space-time, J. Differential Equations 309 (2022), 883–948. 10.1016/j.jde.2021.11.045Search in Google Scholar

[13] Q. Cheng and E. Fan, The Fokas-Lenells equation on the line: global well-posedness with solitons, J. Differential Equations 366 (2023), 320–344. 10.1016/j.jde.2023.04.016Search in Google Scholar

[14] A. Constantin, The Hamiltonian structure of the Camassa-Holm equation, Expo. Math. 15 (1997), 53–85. Search in Google Scholar

[15] A. Constantin, On the scattering problem for the Camassa-Holm equation, Proc. Royal Soc. A. 457 (2001), 953–970. 10.1098/rspa.2000.0701Search in Google Scholar

[16] A. Constantin and J. Escher, Wave breaking for nonlinear nonlocal shallow water equations, Acta Math. 181 (1998), 229–243. 10.1007/BF02392586Search in Google Scholar

[17] A. Constantin and J. Escher, Global existence and blow-up for a shallow water equation, Ann. Scuola Norm. Sup. Pisa Cl. Sci. four 26 (1998), no. 2, 303–328. Search in Google Scholar

[18] A. Constantin, V. Gerdjikov, and R. Ivanov, Inverse scattering transform for the Camassa-Holm equation, Inv. Problems 22 (2006), 2197–2207. 10.1088/0266-5611/22/6/017Search in Google Scholar

[19] A. Constantin and L. Molinet, Global weak solutions for a shallow water equation, Comm. Math. Phys. 211 (2000), 45–61. 10.1007/s002200050801Search in Google Scholar

[20] A. Constantin and L. Molinet, Orbital stability of solitary waves for a shallow water equation, Phys D 157 (2001), 75–89. 10.1016/S0167-2789(01)00298-6Search in Google Scholar

[21] A. Constantin and W. Strauss, Stability of a class of solitary waves in compressible elastic rods, Phys. Lett. A 270 (2000), 140–148. 10.1016/S0375-9601(00)00255-3Search in Google Scholar

[22] A. Constantin and W. Strauss, Stability of peakons, Commun. Pure Appl. Math. 153 (2000), 603–610. 10.1002/(SICI)1097-0312(200005)53:5<603::AID-CPA3>3.0.CO;2-LSearch in Google Scholar

[23] H. Dai, Model equations for nonlinear dispersive waves in a compressible Mooney-Rivlin rod, Acta Mechanica 127 (1998), 193–207. 10.1007/BF01170373Search in Google Scholar

[24] R. Danchin, A few remarks on the Camassa-Holm equation, Differential Integral Equations 14 (2001), no. 8, 953–988. 10.57262/die/1356123175Search in Google Scholar

[25] R. Danchin, A note on well-posedness for Camassa-Holm equation, J. Differential Equations 192 (2003), no. 2, 429–444. 10.1016/S0022-0396(03)00096-2Search in Google Scholar

[26] P. Deift and X. Zhou, A steepest descent method for oscillatory Riemann-Hilbert problems, Ann. Math. 137 (1993), 295–368. 10.2307/2946540Search in Google Scholar

[27] P. Deift and X. Zhou, Long-time asymptotics for solutions of the NLS equation with initial data in a weighted Sobolev space, Comm. Pure Appl. Math. 56 (2003), 1029–1077. 10.1002/cpa.3034Search in Google Scholar

[28] A. S. Fokas, On a class of physically important integrable equations, Phys D Nonlinear Phenomena 87 (1995), 145–150. 10.1016/0167-2789(95)00133-OSearch in Google Scholar

[29] A. S. Fokas and A. A. Himonas, Well-posedness of an integrable generalization of the nonlinear Schrödinger equation on the circle, Lett. Math. Phys. 96 (2011), 169–189. 10.1007/s11005-011-0488-7Search in Google Scholar

[30] B. Fuchssteiner and A. Fokas, Symplectic structures, their Backlund transforms and hereditary symmetries, Phys. D. 4 (1981), 47–66. 10.1016/0167-2789(81)90004-XSearch in Google Scholar

[31] Z. Guo, Local well-posedness and a priori bounds for the modified Benjamin-Ono equation, Adv. Differential Equations 16 (2011), no. 11–12, 1087–1137. 10.57262/ade/1355703113Search in Google Scholar

[32] Z. Guo, X. Liu, L. Molinet, and Z. Yin, Ill-posedness of the Camassa-Holm and related equations in the critical space, J. Differential Equations 266 (2019), no. 2–3, 1698–1707. 10.1016/j.jde.2018.08.013Search in Google Scholar

[33] Z. Guo and Y. Wu, Global well-posedness for the derivative nonlinear Schrödinger equation in H1∕2(R), Discrete Contin. Dyn. Syst. 37 (2017), no. 1, 257–264. 10.3934/dcds.2017010Search in Google Scholar

[34] B. Harrop-Griffiths, R. Killip, M. Ntekoume and M. Vişan, Global well-posedness for the derivative nonlinear Schrödinger equation in L2(R), 2022, arXiv:2204.12548. Search in Google Scholar

[35] J. He, S. Xu, and K. Porsezian, Rogue waves of the Fokas-Lenells equation, J. Phys. Soc. Japan. 81 (2012), Paper no. 124007, 4 pp. 10.1143/JPSJ.81.124007Search in Google Scholar

[36] S. Herr, On the Cauchy problem for the derivative nonlinear Schrödinger equation with periodic boundary condition, Int. Math. Res. Not. 33 (2006), 1–33. 10.1155/IMRN/2006/96763Search in Google Scholar

[37] D. Hilbert, Grundzüge einer allgemeinen Theorie der linearen Integralgleichungen (Erste Mitteilung), Nachr. Ges. Wiss. Göttingen (1904), 49–91. Search in Google Scholar

[38] H. Holden and X. Raynaud, Global conservative solutions of the Camassa-Holm equation-a Lagrangian point of view, Comm. Partial Differential Equations 32 (2007), no. 10–12, 1511–1549. 10.1080/03605300601088674Search in Google Scholar

[39] J. K. Hunter and R. Saxton, Dynamics of director fields, SIAM J. Appl. Math. 51 (1991), 1498–1521. 10.1137/0151075Search in Google Scholar

[40] R. S. Johnson, On solutions of the Camassa-Holm equation, R. Soc. Lond. Proc. Ser. A Math. Phys. Eng. Sci. 459 (2003), 1687–1708. 10.1098/rspa.2002.1078Search in Google Scholar

[41] D. J. Kaup and A. C. Newell, An exact solution for a derivative nonlinear Schrödinger equation, J. Math. Phys. 28 (1978), 798–801. 10.1063/1.523737Search in Google Scholar

[42] R. Killip, T. Laurens and M. Vişan, Sharp well-posedness for the Benjamin-Ono equation, 2023, arXiv 2304.00124. Search in Google Scholar

[43] R. Killip, M. Vişan, and X. Zhang, Low regularity conservation laws for integrable PDE, Geom. Funct. Anal. 28 (2018), 1062–1090. 10.1007/s00039-018-0444-0Search in Google Scholar

[44] R. Killip and M. Vişan, KdV is well-posed in H−1, Ann. Math. 190 (2019), no. 1, 249–305. 10.4007/annals.2019.190.1.4Search in Google Scholar

[45] F. Klaus and R. Schippa, A priori estimates for the derivative nonlinear Schrödinger equation, Funkcial Ekvac. 65 (2022), no. 3, 329–346. 10.1619/fesi.65.329Search in Google Scholar

[46] H. Koch and D. Tataru, Conserved energies for the cubic nonlinear Schrödinger equation in one dimension, Duke Math. J. 167 (2018), no. 1, 3207–3313. 10.1215/00127094-2018-0033Search in Google Scholar

[47] J. Li and Z. Yin, Remarks on the well-posedness of Camassa-Holm type equations in Besov spaces, J. Differential Equations 261 (2016), no. 11, 6125–6143. 10.1016/j.jde.2016.08.031Search in Google Scholar

[48] Y. Li and P. J. Olver, Well-posedness and blow-up solutions for an integrable nonlinearly dispersive model wave equation, J. Differential Equations 162 (2000), no. 1, 27–63. 10.1006/jdeq.1999.3683Search in Google Scholar

[49] J. Lenells, The scattering approach for the Camassa-Holm equation, J. Nonlinear Math. Phys. 9 (2003), no. 4, 389–393. 10.2991/jnmp.2002.9.4.2Search in Google Scholar

[50] J. Lenells, Exactly solvable model for nonlinear pulse propagation in optical fibers, Stud. Appl. Math. 123 (2009), 215–232. 10.1111/j.1467-9590.2009.00454.xSearch in Google Scholar

[51] J. Lenells, Dressing for a novel integrable generalization of the nonlinear Schrödinger equation, J. Nonlinear Sci. 20 (2010), 709–722. 10.1007/s00332-010-9070-1Search in Google Scholar

[52] J. Lenells and A. S. Fokas, An integrable generalization of the nonlinear Schrödinger equation on the half-line and solitons, Inverse Problems 25 (2008), 115006. 10.1088/0266-5611/25/11/115006Search in Google Scholar

[53] J. Lenells and A. S. Fokas, On a novel integrable generalization of the nonlinear Schrödinger equation, Nonlinearity 22 (2009), 11–27. 10.1088/0951-7715/22/1/002Search in Google Scholar

[54] Y. Matsuno, A direct method of solution for the Fokas-Lenells derivative nonlinear Schrödinger equation: I. Bright soliton solutions, 45 (2012), 235202, 19. 10.1088/1751-8113/45/23/235202Search in Google Scholar

[55] Y. Matsuno, A direct method of solution for the Fokas-Lenells derivative nonlinear Schrödinger equation: II. Dark soliton solutions, J. Phys. A 45 (2012), 475202, 31. 10.1088/1751-8113/45/47/475202Search in Google Scholar

[56] H. P. McKean, Breakdown of a shallow water equation, Asian J. Math. 2 (1998), no. 4, 867–874. 10.4310/AJM.1998.v2.n4.a10Search in Google Scholar

[57] A. B. Monvel, A. Kostenko, D. Shepelsky, and G. Teschl, Long-time asymptotics for the Camassa-Holm equation, SIAM J. Math. Anal. 41 (2009), 1559–1588. 10.1137/090748500Search in Google Scholar

[58] B. Moon, Single peaked traveling wave solutions to a generalized μ-Novikov equation, Adv. Nonlinear Anal. 10 (2012), no. 1, 66–75. 10.1515/anona-2020-0106Search in Google Scholar

[59] R. Mosincat, Global well-posedness of the derivative nonlinear Schrödinger equation with periodic boundary condition in H1∕2, J. Differential Equations 263 (2017), no. 8, 4658–4722. 10.1016/j.jde.2017.05.026Search in Google Scholar

[60] R. Mosincat and T. Oh, A remark on global well-posedness of the derivative nonlinear Schrödinger equation on the circle, C. R. Math. Acad. Sci. Paris, 353 (2015), no. 9, 837–841. 10.1016/j.crma.2015.06.015Search in Google Scholar

[61] T. Oh and Y. Wang, Global well-posedness of the one-dimensional cubic nonlinear Schrödinger equation in almost critical spaces, J. Differential Equations 269 (2020), no. 1, 612–640. 10.1016/j.jde.2019.12.017Search in Google Scholar

[62] C. Qu, X. Liu, and Y. Liu, Stability of peakons for an integrable modified Camassa-Holm equation with cubic nonlinearity, Comm. Math. Phys. 322 (2013), no. 3, 967–997. 10.1007/s00220-013-1749-3Search in Google Scholar

[63] A. Rybkin, Regularized perturbation determinants and KdV conservation laws for irregular initial profiles, Topics Operator Theory 203 (2010), 427–444. 10.1007/978-3-0346-0161-0_17Search in Google Scholar

[64] R. Schippa, On a priori estimates and existence of periodic solutions to the modified Benjamin-Ono equation below H1∕2(T), 2017, arXiv:1704.07174. Search in Google Scholar

[65] B. Simon, Trace ideals and their applications, Second edition, Mathematical Surveys and Monographs, vol. 120, A.M.S., Providence, RI, 2005, viii+150pp. Search in Google Scholar

[66] B. Simon, Operator theory, A Comprehensive Course in Analysis, Part 4, American Mathematical Society, Providence, RI, 2015. Search in Google Scholar

[67] H. Takaoka, Well-posedness for the one-dimensional nonlinear Schrödinger equation with the derivative nonlinearity, Adv. Differential Equations 4 (1999), no. 4, 561–580. 10.57262/ade/1366031032Search in Google Scholar

[68] B. Talbut, Low regularity conservation laws for the Benjamin-Ono equation, Math. Res. Lett. 28 (2021), no. 3, 889–905. 10.4310/MRL.2021.v28.n3.a11Search in Google Scholar

[69] X. Tang and G. Xu, Microscopic conservation laws for the derivative nonlinear Schrödinger equation, 2020, arXiv:2012.04805. 10.1007/s11005-021-01478-ySearch in Google Scholar

[70] Z. Wang, L. He, Z. Qin, R. Grimshaw, and G. Mu, High-order rogue waves and their dynamics of the Fokas-Lenells equation revisited: a variable separation technique, Nonlinear Dyn. 98 (2019), 2067–2077. 10.1007/s11071-019-05308-8Search in Google Scholar

[71] Y. Wu, Global well-posedness for the nonlinear Schrödinger equation with derivative in energy space, Anal. PDE 6 (2013), no. 8, 1989–2002. 10.2140/apde.2013.6.1989Search in Google Scholar

[72] Y. Wu, Global well-posedness on the derivative nonlinear Schrödinger equation, Anal. PDE 8 (2015), no. 5, 1101–1112. 10.2140/apde.2015.8.1101Search in Google Scholar

[73] Y. Xiao, E. Fan, and J. Xu, The Fokas-Lenells equation on the finite interval, Acta Math. Sci. Ser. B (Engl. Ed.) 37 (2017), 852–876. 10.1016/S0252-9602(17)30041-3Search in Google Scholar

[74] Z. Xin and P. Zhang, On the weak solutions to a shallow water equations, Commun. Pure Appl. Math. 53 (2000), 1411–1433. 10.1002/1097-0312(200011)53:11<1411::AID-CPA4>3.0.CO;2-5Search in Google Scholar

[75] J. Xu and E. Fan, Long-time asymptotics for the Fokas-Lenells equation with decaying initial value problem: Without solitons, J. Differential Equations 259 (2015), 1098–1148. 10.1016/j.jde.2015.02.046Search in Google Scholar

[76] R. Xu and Y. Yan, Low regularity of solutions to the Rotation-Camassa-Holm type equation with the Coriolis effect, Discrete Contin. Dyn. Syst. 40 (2020), no. 11, 6527–6547. 10.3934/dcds.2020288Search in Google Scholar

[77] R. Xu and Y. Yan, Local well-posedness and decay for some generalized shallow water equations, J. Differential Equations 367 (2023), no. 15, 689–728. 10.1016/j.jde.2023.05.026Search in Google Scholar

[78] S. Xu, J. He, Y. Cheng, and K. Porseizan, The n-order rogue waves of Fokas-Lenells equation, Math. Meth. Appl. Sci. 38 (2015), 1106–1126. 10.1002/mma.3133Search in Google Scholar

[79] D. R. Yafaev, Mathematical scattering theory, General theory, Translated from Russian by J. R. Schulenberger, Translation of Mathematical Monographs, vol. 105, A.M.S., Providence, RI, 1992, x+341pp. Search in Google Scholar

[80] Y. Zhao and E. Fan, Inverse scattering transformation for the Fokas-Lenells equation with nonzero boundary conditions, J. Nonl. Math. Phys. 28 (2021), 38–52. 10.2991/jnmp.k.200922.003Search in Google Scholar

[81] P. Zhao, E. Fan, and Y. Hou, Algebro-geometric solutions and their reductions for the Fokas-Lenells hierarchy, J. Nonl. Math. Phys. 20 (2013), 355–393. 10.1080/14029251.2013.854094Search in Google Scholar

Received: 2023-09-13
Revised: 2024-01-19
Accepted: 2024-04-04
Published Online: 2024-06-08

© 2024 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Regular Articles
  2. Existence of normalized peak solutions for a coupled nonlinear Schrödinger system
  3. Orbital stability of the trains of peaked solitary waves for the modified Camassa-Holm-Novikov equation
  4. Global boundedness in a two-dimensional chemotaxis system with nonlinear diffusion and singular sensitivity
  5. Existence and uniqueness of solution for a singular elliptic differential equation
  6. The bounded variation capacity and Sobolev-type inequalities on Dirichlet spaces
  7. Vanishing and blow-up solutions to a class of nonlinear complex differential equations near the singular point
  8. Existence, uniqueness, localization and minimization property of positive solutions for non-local problems involving discontinuous Kirchhoff functions
  9. Concentration phenomena for a fractional relativistic Schrödinger equation with critical growth
  10. Beyond the classical strong maximum principle: Sign-changing forcing term and flat solutions
  11. Monotonicity of solutions for parabolic equations involving nonlocal Monge-Ampère operator
  12. On a nonlinear Robin problem with an absorption term on the boundary and L1 data
  13. Multiple solutions for the quasilinear Choquard equation with Berestycki-Lions-type nonlinearities
  14. Gradient estimates for a class of higher-order elliptic equations of p-growth over a nonsmooth domain
  15. Global existence and decay estimates of the classical solution to the compressible Navier-Stokes-Smoluchowski equations in ℝ3
  16. Global existence and finite-time blowup for a mixed pseudo-parabolic r(x)-Laplacian equation
  17. Multiple positive solutions for a class of concave-convex Schrödinger-Poisson-Slater equations with critical exponent
  18. A minimization problem with free boundary for p-Laplacian weakly coupled system
  19. Multiplicity of semiclassical solutions for a class of nonlinear Hamiltonian elliptic system
  20. k-convex solutions for multiparameter Dirichlet systems with k-Hessian operator and Lane-Emden type nonlinearities
  21. Infinitely many solutions for Hamiltonian system with critical growth
  22. On optimal control in a nonlinear interface problem described by hemivariational inequalities
  23. Variational–hemivariational system for contaminant convection–reaction–diffusion model of recovered fracturing fluid
  24. Potential and monotone homeomorphisms in Banach spaces
  25. Critical fractional Schrödinger-Poisson systems with lower perturbations: the existence and concentration behavior of ground state solutions
  26. Supercritical Hénon-type equation with a forcing term
  27. Concentration of blow-up solutions for the Gross-Pitaveskii equation
  28. Mountain-pass-type solutions for Schrödinger equations in R2 with unbounded or vanishing potentials and critical exponential growth nonlinearities
  29. Regularity for critical fractional Choquard equation with singular potential and its applications
  30. Double phase anisotropic variational problems involving critical growth
  31. Normalized solutions of NLS equations with mixed nonlocal nonlinearities
  32. Nontrivial solutions for resonance quasilinear elliptic systems
  33. Standing waves for Choquard equation with noncritical rotation
  34. Low regularity conservation laws for Fokas-Lenells equation and Camassa-Holm equation
  35. Modified quasilinear equations with strongly singular and critical exponential nonlinearity
  36. Limit profiles and the existence of bound-states in exterior domains for fractional Choquard equations with critical exponent
  37. Nests of limit cycles in quadratic systems
  38. Regularity of minimizers for double phase functionals of borderline case with variable exponents
  39. Boundedness, existence and uniqueness results for coupled gradient dependent elliptic systems with nonlinear boundary condition
  40. Ground state solutions for magnetic Schrödinger equations with polynomial growth
  41. Nonoccurrence of Lavrentiev gap for a class of functionals with nonstandard growth
  42. Dynamics for wave equations connected in parallel with nonlinear localized damping
  43. Boussinesq's equation for water waves: Asymptotics in Sector I
  44. Optimal global second-order regularity and improved integrability for parabolic equations with variable growth
  45. Normalized solutions of Schrödinger equations involving Moser-Trudinger critical growth
  46. Normalized solutions for the double-phase problem with nonlocal reaction
  47. Normalized solutions for Sobolev critical fractional Schrödinger equation
  48. Well-posedness of Cauchy problem of fractional drift diffusion system in non-critical spaces with power-law nonlinearity
  49. Improved results on planar Klein-Gordon-Maxwell system with critical exponential growth
  50. Existence and multiplicity of solutions for a new p(x)-Kirchhoff equation
  51. Quasiconvex bulk and surface energies: C1,α regularity
  52. Time decay estimates of solutions to a two-phase flow model in the whole space
  53. Bounds for the sum of the first k-eigenvalues of Dirichlet problem with logarithmic order of Klein-Gordon operators
  54. The properties of a new fractional g-Laplacian Monge-Ampère operator and its applications
  55. A fractional profile decomposition and its application to Kirchhoff-type fractional problems with prescribed mass
  56. Cauchy problem for a non-Newtonian filtration equation with slowly decaying volumetric moisture content
  57. Multiplicity of normalized semi-classical states for a class of nonlinear Choquard equations
  58. The second Yamabe invariant with boundary
  59. Peaked solitary waves and shock waves of the Degasperis-Procesi-Kadomtsev-Petviashvili equation
  60. Normalized solutions for the Choquard equations with critical nonlinearities
  61. Boundedness and long-time behavior in a parabolic-elliptic system arising from biological transport networks
  62. Initial boundary value problem and exponential stability for the planar magnetohydrodynamics equations with temperature-dependent viscosity
  63. Nonexistence of mean curvature flow solitons with polynomial volume growth immersed in certain semi-Riemannian warped products
  64. Analysis of a vector-borne disease model with vector-bias mechanism in advective heterogeneous environment
  65. Normalized solutions for the Kirchhoff equation with combined nonlinearities in ℝ4
  66. Nonexistence of the compressible Euler equations with space-dependent damping in high dimensions
  67. Multiplicity of solutions for a nonhomogeneous quasilinear elliptic equation with concave-convex nonlinearities
  68. On semilinear inequalities involving the Dunkl Laplacian and an inverse-square potential outside a ball
  69. Besov regularity for the elliptic p-harmonic equations in the non-quadratic case
  70. Uniqueness and nondegeneracy of ground states for Δ u + u = ( I α u 2 ) u in R 3 when α is close to 2
  71. Existence of solutions for a class of quasilinear Schrödinger equations with Choquard-type nonlinearity
  72. Weighted Hardy-Adams inequality on unit ball of any even dimension
  73. Existence and regularity for a p-Laplacian problem in ℝN with singular, convective, and critical reaction
  74. Temporal periodic solutions of non-isentropic compressible Euler equations with geometric effects
  75. Nodal solutions for a zero-mass Chern-Simons-Schrödinger equation
  76. The modified quasi-boundary-value method for an ill-posed generalized elliptic problem
  77. Nonlocal heat equations with generalized fractional Laplacian
  78. Choquard equations with recurrent potentials
  79. Local and global well-posedness of the Maxwell-Bloch system of equations with inhomogeneous broadening
  80. The Riemann problem for two-layer shallow water equations with bottom topography
  81. Global stability and asymptotic profiles of a partially degenerate reaction diffusion Cholera model with asymptomatic individuals
  82. On Kirchhoff-Schrödinger-Poisson-type systems with singular and critical nonlinearity
  83. Energy-variational solutions for viscoelastic fluid models
  84. Stability on 3D Boussinesq system with mixed partial dissipation
  85. Existence and multiplicity results for non-autonomous second-order Hamiltonian systems
  86. Erratum
  87. Erratum to “Normalized solutions for the Kirchhoff equation with combined nonlinearities in ℝ4
Downloaded on 13.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/anona-2024-0014/html
Scroll to top button