• Aucun résultat trouvé

Error estimates of energy regularization for the logarithmic Schrodinger equation

N/A
N/A
Protected

Academic year: 2021

Partager "Error estimates of energy regularization for the logarithmic Schrodinger equation"

Copied!
32
0
0

Texte intégral

(1)

HAL Id: hal-02817115

https://hal.archives-ouvertes.fr/hal-02817115v2

Preprint submitted on 5 Sep 2021

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Error estimates of local energy regularization for the logarithmic Schrodinger equation

Weizhu Bao, Rémi Carles, Chunmei Su, Qinglin Tang

To cite this version:

Weizhu Bao, Rémi Carles, Chunmei Su, Qinglin Tang. Error estimates of local energy regularization

for the logarithmic Schrodinger equation. 2021. �hal-02817115v2�

(2)

ERROR ESTIMATES OF LOCAL ENERGY REGULARIZATION FOR THE LOGARITHMIC SCHR ¨ ODINGER EQUATION

WEIZHU BAO, R´ EMI CARLES, CHUNMEI SU, AND QINGLIN TANG

Abstract. The logarithmic nonlinearity has been used in many partial differ- ential equations (PDEs) for modeling problems in various applications. Due to the singularity of the logarithmic function, it introduces tremendous difficulties in establishing mathematical theories, as well as in designing and analyzing nu- merical methods for PDEs with such nonlinearity. Here we take the logarithmic Schr¨ odinger equation (LogSE) as a prototype model. Instead of regularizing f(ρ) = ln ρ in the LogSE directly and globally as being done in the literature, we propose a local energy regularization (LER) for the LogSE by first regular- izing F(ρ) = ρln ρ − ρ locally near ρ = 0

+

with a polynomial approximation in the energy functional of the LogSE and then obtaining an energy regular- ized logarithmic Schr¨ odinger equation (ERLogSE) via energy variation. Linear convergence is established between the solutions of ERLogSE and LogSE in terms of a small regularization parameter 0 < ε ≪ 1. Moreover, the conserved energy of the ERLogSE converges to that of LogSE quadratically, which sig- nificantly improves the linear convergence rate of the regularization method in the literature. Error estimates are also presented for solving the ERLogSE by using Lie-Trotter splitting integrators. Numerical results are reported to confirm our error estimates of the LER and of the time-splitting integrators for the ERLogSE. Finally our results suggest that the LER performs better than regularizing the logarithmic nonlinearity in the LogSE directly.

1. Introduction

The logarithmic nonlinearity appears in physical models from many fields. For example, the logarithmic nonlinearity is introduced in quantum mechanics or quan- tum optics, where a logarithmic Schr¨odinger equation (LogSE) is considered (e.g.

[14–16, 44]),

i∂

t

u = − ∆u + λ u ln | u | 2 , λ ∈ R ;

in oceanography and in fluid dynamics, with a logarithmic Korteweg-de Vries (KdV) equation or a logarithmic Kadomtsev-Petviashvili (KP) equation (e.g. [39, 50, 51]);

in quantum field theory and in inflation cosmology, via a logarithmic Klein-Gordon equation (e.g. [12, 35, 49]); or in material sciences, by the introduction of a Cahn- Hilliard (CH) equation with logarithmic potentials (e.g. [24, 28, 33]). Recently, the

2020 Mathematics Subject Classification. 22E46, 53C35, 57S235Q40, 35Q55, 65M15, 81Q05.

Key words and phrases. Logarithmic Schr¨ odinger equation; logarithmic nonlinearity; energy regularization; error estimates; convergence rate; Lie-Trotter splitting.

This work was partially supported by the Ministry of Education of Singapore grant R-146-000- 296-112 (MOE2019-T2-1-063) (W. Bao), Rennes M´ etropole through its AIS program (R. Carles), the Alexander von Humboldt Foundation (C. Su), the Institutional Research Fund from Sichuan University (No. 2020SCUNL110) and the National Natural Science Foundation of China (No.

11971335) (Q. Tang).

1

(3)

heat equation with a logarithmic nonlinearity has been investigated mathematically [1, 22].

In the context of quantum mechanics, the logarithmic nonlinearity was selected by assuming the separability of noninteracting subsystems property (cf. [14]). This means that a solution of the nonlinear equation for the whole system can be con- structed, as in the linear theory, by taking the product of two arbitrary solutions of the nonlinear equations for the subsystems. In other words, no correlations are introduced for noninteracting subsystems. As for the physical reality, robust physical grounds have been found for the application of equations with logarithmic nonlinearity. For instance, it was found in the stochastic formulation of quantum mechanics [45, 48] that the logarithmic nonlinear term originates naturally from an internal stochastic force due to quantum fluctuations. Such kind of nonlin- earity also appears naturally in inflation cosmology and in supersymmetric field theories [11, 30].

Remarkably enough for a nonlinear PDE, many explicit solutions are available for the logarithmic mechanics (see e.g. [14, 43]). For example, the logarithmic KdV equation, the logarithmic KP equation, the logarithmic Klein-Gordon equation give Gaussons: solitary wave solutions with Gaussian shapes [50, 51]. In the case of LogSE (see [17,31]), or the heat equation [1], every initial Gaussian function evolves as a Gaussian: solving the corresponding nonlinear PDE is equivalent to solving ordinary differential equations (involving the purely time dependent parameters of the Gaussian). However we emphasize that this is not so in the case of, e.g., the logarithmic KdV equation, the logarithmic KP equation, or the logarithmic Klein- Gordon equation. This can be directly seen by trying to plug time dependent Gaussian functions into these equations. Note that this distinction between various PDEs regarding the propagation of Gaussian functions is the same as at the linear level.

The well-posedness of the Cauchy problem for logarithmic equations is not trivial since the logarithmic nonlinearity is not locally Lipschitz continuous, due to the singularity of the logarithm at the origin. Existence was proved by compactness argument based on regularization of the nonlinearity, for the CH equation with a logarithmic potential [29] and the LogSE [18]. Uniqueness is also a challenging question, settled in the case of LogSE thanks to a surprising inequality discovered in [20], recalled in Lemma 2.1 below.

The singularity of the logarithmic nonlinearity also makes it very challenging to design and analyze numerical schemes. There have been extensive numerical works for the CH equation with a logarithmic Flory Huggins energy potential [23, 25, 34, 40, 41, 52]. Specifically, a regularized energy functional was adopted for the CH equation with a logarithmic free energy [25,52]. A regularization of the logarithmic nonlinearity was introduced and analyzed in [4, 5] in the case LogSE, see also [46].

In this paper, we introduce and analyze numerical methods for logarithmic equa- tions via a local energy regularization. We consider the LogSE as an example; the regularization can be extended to other logarithmic equations. The LogSE which arises in a model of nonlinear wave mechanics reads (cf. [14]),

(1.1)

( i∂

t

u(x, t) = − ∆u(x, t) + λ u(x, t) f ( | u(x, t) | 2 ), x ∈ Ω, t > 0,

u(x, 0) = u 0 (x), x ∈ Ω,

(4)

where t and x ∈ R

d

(d = 1, 2, 3) represent the temporal and spatial coordi- nates, respectively, λ ∈ R \{ 0 } measures the force of the nonlinear interaction, u := u( x , t) ∈ C is the dimensionless wave function, and

(1.2) f (ρ) = ln ρ, ρ > 0, with ρ = | u | 2 .

The spatial domain is either Ω = R

d

, or Ω ⊂ R

d

bounded with Lipschitz continuous boundary; in the latter case the equation is subject to homogeneous Dirichlet or periodic boundary conditions. This model has been widely applied in quantum mechanics, nuclear physics, geophysics, open quantum systems and Bose-Einstein condensation, see e.g. [3, 26, 37, 38, 53]. We choose to consider positive time only merely to simplify the presentation, since (1.1) is time reversible. Formally, the flow of (1.1) enjoys two important conservations. The mass, defined as

(1.3) N (t) := N (u( · , t)) = k u k 2 = Z

Ω | u(x, t) | 2 dx ≡ N(u 0 ), t ≥ 0, and the energy, defined as

E(t) : = E(u( · , t)) = Z

|∇ u(x, t) | 2 dx + λF ( | u(x, t) | 2 ) dx

≡ Z

|∇ u 0 ( x ) | 2 + λF ( | u 0 ( x ) | 2 )

d x = E(u 0 ), t ≥ 0, (1.4)

where

(1.5) F (ρ) = Z

ρ

0

f (s)ds = Z

ρ

0

ln s ds = ρ ln ρ − ρ, ρ ≥ 0.

The total angular momentum is also conserved, an identity that we do not use in the present paper. For the Cauchy problem (1.1) in a suitable functional framework, we refer to [17, 20, 36]. For stability properties of standing waves for (1.1), we refer to [2, 18, 21]. For the analysis of breathers and the existence of multisolitons, see [31, 32].

In order to avoid numerical blow-up of the logarithmic nonlinearity at the ori- gin, two models of regularized logarithmic Schr¨odinger equation (RLogSE) were proposed in [5], involving a direct regularization of f in (1.2), relying on a small regularized parameter 0 < ε ≪ 1,

(1.6)

( i∂

t

u

ε

(x, t) = − ∆u

ε

(x, t) + λ u

ε

(x, t) f e

ε

( | u

ε

(x, t) | ) 2 ), x ∈ Ω, t > 0, u

ε

(x, 0) = u 0 (x), x ∈ Ω,

and (1.7)

( i∂

t

u

ε

(x, t) = − ∆u

ε

(x, t) + λ u

ε

(x, t) f b

ε

( | u

ε

(x, t) | 2 )), x ∈ Ω, t > 0, u

ε

(x, 0) = u 0 (x), x ∈ Ω.

Here, f e

ε

(ρ) and f b

ε

(ρ) are two types of regularization for f (ρ), given by

(1.8) f e

ε

(ρ) = 2 ln(ε + √ ρ), f b

ε

(ρ) = ln(ε 2 + ρ), ρ ≥ 0, with ρ = | u

ε

| 2 . Again, the RLogSEs (1.6) and (1.7) conserve the mass (1.3) with u = u

ε

, as well as the energies

(1.9) E e

ε

(t) := E e

ε

(u

ε

( · , t)) = Z

h

|∇ u

ε

(x, t) | 2 dx + λ F e

ε

( | u

ε

(x, t) | 2 ) i

dx ≡ E e

ε

(u 0 ),

(5)

and (1.10)

E b

ε

(t) := E b

ε

(u

ε

( · , t)) = Z

h |∇ u

ε

(x, t) | 2 dx + λ F b

ε

( | u

ε

(x, t) | 2 ) i

dx ≡ E b

ε

(u 0 ), respectively, with, for ρ ≥ 0,

F e

ε

(ρ) = Z

ρ

0

f e

ε

(s)ds = 2ρ ln(ε + √ ρ) + 2ε √ ρ − ρ − 2ε 2 ln(1 + √ ρ/ε), F b

ε

(ρ) =

Z

ρ

0

f b

ε

(s)ds = (ε 2 + ρ) ln(ε 2 + ρ) − ρ − 2ε 2 ln ε.

(1.11)

The idea of this regularization is that the function ρ 7→ ln ρ causes no (analytical or numerical) problem for large values of ρ, but is singular at ρ = 0. A linear conver- gence was established between the solutions of the LogSE (1.1) and the regularized model (1.6) or (1.7) for bounded Ω in terms of the small regularization parameter 0 < ε ≪ 1, i.e.,

sup

t∈

[0,T] k u

ε

(t) − u(t) k

L2

(Ω) = O(ε), ∀ T > 0.

Applying this regularized model, a semi-implicit finite difference method (FDM) and a time-splitting method were proposed and analyzed for the LogSE (1.6) in [5]

and [4] respectively. The above regularization saturates the nonlinearity in the region { ρ < ε 2 } (where ρ = | u

ε

| 2 ), but of course has also some (smaller) effect in the other region { ρ > ε 2 } , i.e., it regularizes f (ρ) = ln ρ globally.

Energy regularization is a method which has been adapted in different fields for dealing with singularity and/or roughness: in materials science, for establishing the well-posedness of the Cauchy problem for the CH equation with a logarithmic potential [29], and for treating strongly anisotropic surface energy [7,42]; in mathe- matical physics, for the well-posedness of the LogSE [18]; in scientific computing, for designing regularized numerical methods in the presence of singularities [9, 25, 52].

The main goal of this paper is to present a local energy regularization (LER) for the LogSE (1.1). We regularize the interaction energy density F(ρ) only locally in the region { ρ < ε 2 } by a sequence of polynomials, and keep it unchanged in { ρ > ε 2 } . The choice of the regularized interaction energy density F

nε

is prescribed by the regularity n imposed at this step, involving the matching conditions at { ρ = ε 2 } . We then obtain a sequence of energy regularized logarithmic Schr¨odinger equations (ERLogSEs), from the regularized energy functional density F

nε

, via energy varia- tion. Unlike in [25, 52], where the interaction energy density F (ρ) is approximated by a second order polynomial near the origin, here we present a systematic way to regularize the interaction energy density near the origin, i.e. locally, by a se- quence of polynomials such that the order of regularity n of the overall regularized interaction energy density is arbitrary. We establish convergence rates between the solutions of ERLogSEs and LogSE in terms of the small regularized parameter 0 < ε ≪ 1. In addition, we also prove error estimates of numerical approximations of ERLogSEs by using time-splitting integrators.

The rest of this paper is organized as follows. In Section 2, we introduce a

sequence of regularization F

nε

for the logarithmic potential. A regularized model

is derived and analyzed in Section 3 via the LER of the LogSE. Some numerical

methods are proposed and analyzed in Section 4. In Section 5, we present numerical

experiments. Throughout the paper, we adopt the standard L 2 -based Sobolev

(6)

spaces as well as the corresponding norms, and denote by C a generic positive constant independent of ε, the time step τ and the function u, and by C(c) a generic positive constant depending on c.

2. Local regularization for F(ρ) = ρ ln ρ − ρ

We consider a local regularization starting from an approximation to the inter- action energy density F (ρ) in (1.5) (and thus in (1.4)).

2.1. A sequence of local regularization. In order to make a comparison with the former global regularization (1.6), we again distinguish the regions { ρ > ε 2 } and { ρ < ε 2 } . Instead of saturating the nonlinearity in the second region, we regularize it locally as follows. For an arbitrary integer n ≥ 2, we approximate F (ρ) by a piecewise smooth function which is polynomial near the origin,

(2.1) F

nε

(ρ) = F(ρ)χ

{ρ≥ε2}

+ P

n+1ε

(ρ)χ

{ρ<ε2}

, n ≥ 2,

where 0 < ε ≪ 1 is a small regularization parameter, χ

A

is the characteristic function of the set A, and P

n+1ε

is a polynomial of degree n + 1. We demand F

nε

∈ C

n

([0, + ∞ )) and F

nε

(0) = F(0) = 0 (this allows the regularized energy to be well-defined on the whole space). The above conditions determine P

n+1ε

, as we now check. Since P

n+1ε

(0) = 0, write

(2.2) P

n+1ε

(ρ) = ρ Q

εn

(ρ),

with Q

εn

a polynomial of degree n. Correspondingly, denote F (ρ) = ρ Q(ρ) with Q(ρ) = ln ρ − 1. The continuity conditions read

P

n+1ε

2 ) = F(ε 2 ), (P

n+1ε

)

2 ) = F

2 ), . . . , (P

n+1ε

) (n)2 ) = F (n)2 ), which in turn yield

Q

εn

2 ) = Q(ε 2 ), (Q

εn

)

2 ) = Q

2 ), . . . , (Q

εn

) (n)2 ) = Q (n)2 ).

Thus Q

εn

is nothing else but Taylor polynomial of Q of degree n at ρ = ε 2 , i.e., (2.3) Q

εn

(ρ) = Q(ε 2 ) +

X

n

k=1

Q (k)2 )

k! (ρ − ε 2 )

k

= ln ε 2 − 1 − X

n

k=1

1 k

1 − ρ ε 2

k

.

In particular, Taylor’s formula yields (2.4) Q(ρ) − Q

εn

(ρ) =

Z

ρ ε2

Q (n+1) (s) (ρ − s)

n

n! ds =

Z

ρ ε2

(s − ρ)

n

s

n+1

ds.

Plugging (2.3) into (2.2), we get the explicit formula of P

n+1ε

(ρ). We emphasize a formula which will be convenient for convergence results:

(2.5) (Q

εn

)

(ρ) = 1 ε 2

X

n

k=1

1 − ρ ε 2

k−

1

= 1 ρ

1 − 1 − ρ

ε 2

n

, 0 ≤ ρ ≤ ε 2 .

(7)

2.2. Properties of the local regularization functions. Differentiating (2.1) with respect to ρ and noting (2.2), (2.3) and (2.5), we get

(2.6) f

nε

(ρ) = (F

nε

)

(ρ) = ln ρ χ

{ρ≥ε2}

+ q

nε

(ρ)χ

{ρ<ε2}

, ρ ≥ 0, where

q

nε

(ρ) = (P

n+1ε

)

(ρ) = Q

εn

(ρ) + ρ (Q

εn

)

(ρ)

= ln(ε 2 ) − n + 1 n

1 − ρ ε 2

n

n−

1

X

k=1

1 k

1 − ρ ε 2

k

.

Noticing that q

εn

is increasing in [0, ε 2 ], f e

ε

and f b

ε

are increasing on [0, ∞ ), thus all three types of regularization (2.1) and (1.11) preserve the convexity of F . Moreover, as a sequence of local regularization (or approximation) for the semi- smooth function F (ρ) ∈ C 0 ([0, ∞ )) ∩ C

((0, ∞ )), we have F

nε

∈ C

n

([0, + ∞ )) for n ≥ 2, while F e

ε

∈ C 1 ([0, ∞ )) ∩ C

((0, ∞ )) and F b

ε

∈ C

([0, ∞ )). Similarly, as a sequence of local regularization (or approximation) for the logarithmic function f (ρ) = ln ρ ∈ C

((0, ∞ )), we observe that f

nε

∈ C

n

1 ([0, ∞ )) for n ≥ 2, while f b

ε

∈ C

([0, ∞ )) and f e

ε

∈ C 0 ([0, ∞ )) ∩ C

((0, ∞ )).

Recall the following lemma, established initially in [20, Lemma 1.1.1].

Lemma 2.1. For z 1 , z 2 ∈ C , we have Im z 1 ln( | z 1 | 2 ) − z 2 ln( | z 2 | 2 )

(z 1 − z 2 ) ≤ 2 | z 1 − z 2 | 2 ,

where Im(z) and z denote the imaginary part and the complex conjugate of z, re- spectively.

Next we highlight some properties of f

nε

.

Lemma 2.2. Let n ≥ 2 and ε > 0. For z 1 , z 2 ∈ C , we have

| f

nε

( | z 1 | 2 ) − f

nε

( | z 2 | 2 ) | ≤ 4n | z 1 − z 2 | max { ε, min {| z 1 | , | z 2 |}} , (2.7)

Im

z 1 f

nε

( | z 1 | 2 ) − z 2 f

nε

( | z 2 | 2 )

(z 1 − z 2 ) ≤ 4n | z 1 − z 2 | 2 , (2.8)

| ρ(f

nε

)

(ρ) | ≤ 3, | √ ρ(f

nε

)

(ρ) | ≤ 2n

ε , | ρ 3/2 (f

nε

)

′′

(ρ) | ≤ 3n 2

2ε , ρ ≥ 0, (2.9)

| f

nε

(ρ) | ≤ max {| ln A | , 2 + ln(nε

2 ) } , ρ ∈ [0, A].

(2.10)

Proof. When | z 1 | , | z 2 | ≥ ε, we have f

nε

( | z 1 | 2 ) − f

nε

( | z 2 | 2 ) = 2 ln

1 + || z 1 | − | z 2 ||

min {| z 1 | , | z 2 |}

≤ 2 | z 1 − z 2 | min {| z 1 | , | z 2 |} . A direct calculation gives

(2.11) (f

nε

)

(ρ) = 1

ρ χ

{ρ≥ε2}

+ n ε 2 1 − ρ

ε 2

n−

1

+ 1 ε 2

n−

1

X

k=0

1 − ρ ε 2

k

!

χ

{ρ<ε2}

.

Thus when | z 1 | < | z 2 | ≤ ε, we have f

nε

( | z 1 | 2 ) − f

nε

( | z 2 | 2 ) =

Z

|z2|2

|z1|2

(f

nε

)

(ρ)dρ

= n ε 2

Z

|z2|2

|z1|2

1 − ρ ε 2

n−

1

dρ + 1 ε 2

n

X

1

k=0

Z

|z2|2

|z1|2

1 − ρ ε 2

k

(8)

≤ n

ε 2 ( | z 2 | 2 − | z 1 | 2 ) + 1 ε 2

n−

1

X

k=0

( | z 2 | 2 − | z 1 | 2 )

= 2n

ε 2 ( | z 2 | 2 − | z 1 | 2 ) ≤ 4n

ε | z 1 − z 2 | .

Another case when | z 2 | < | z 1 | ≤ ε can be established similarly. Supposing, for example, | z 2 | < ε < | z 1 | , denote by z 3 the intersection point of the circle { z ∈ C :

| z | = ε } and the line segment connecting z 1 and z 2 . Combining the inequalities above, we have

f

nε

( | z 1 | 2 ) − f

nε

( | z 2 | 2 ) ≤ | f

nε

( | z 2 | 2 ) − f

nε

( | z 3 | 2 ) | + | ln( | z 1 | 2 ) − ln( | z 3 | 2 ) |

≤ 4n

ε | z 2 − z 3 | + 2

ε | z 1 − z 3 |

≤ 4n

ε ( | z 2 − z 3 | + | z 1 − z 3 | ) = 4n

ε | z 1 − z 2 | , which completes the proof for (2.7).

Noticing that Im

z 1 f

nε

( | z 1 | 2 ) − z 2 f

nε

( | z 2 | 2 )

(z 1 − z 2 )

= − Im(z 1 z 2 )f

nε

( | z 2 | 2 ) − Im(z 1 z 2 )f

nε

( | z 1 | 2 )

= Im(z 1 z 2 )

f

nε

( | z 1 | 2 ) − f

nε

( | z 2 | 2 )

= 1

2i (z 1 z 2 − z 1 z 2 )

f

nε

( | z 1 | 2 ) − f

nε

( | z 2 | 2 ) , and

| z 1 z 2 − z 1 z 2 | = | z 2 (z 1 − z 2 ) + z 2 (z 2 − z 1 ) | ≤ 2 | z 2 | | z 1 − z 2 | ,

| z 1 z 2 − z 1 z 2 | = | z 1 (z 2 − z 1 ) + z 1 (z 1 − z 2 ) | ≤ 2 | z 1 | | z 1 − z 2 | , which implies

| z 1 z 2 − z 1 z 2 | ≤ 2 min {| z 1 | , | z 2 |} | z 1 − z 2 | , one can conclude (2.8) by applying (2.7).

It follows from (2.11) that

g(ρ) = ρ(f

nε

)

(ρ) = χ

{ρ≥ε2}

+ nρ ε 2 1 − ρ

ε 2

n−

1

+ ρ ε 2

n

X

1

k=0

1 − ρ ε 2

k

!

χ

{ρ<ε2}

= χ

{ρ≥ε2}

+ nρ ε 2 1 − ρ

ε 2

n−

1

+ 1 − 1 − ρ ε 2

n

χ

{ρ<ε2}

, which gives that

g

(ρ)χ

{ρ<ε2}

= n ε 2 1 − ρ

ε 2

n−

2

2 − (n + 1)ρ ε 2

.

This leads to

| ρ(f

nε

)

(ρ) | = g(ρ) ≤ max { 1, g 2ε 2

n + 1

} ≤ 1 + 2n n + 1 ≤ 3,

which completes the proof for the first inequality in (2.9). Finally it follows from (2.11) that

√ ρ(f

nε

)

(ρ) = 1

√ ρ χ

{ρ≥ε2}

+

√ ρ

ε 2 n 1 − ρ ε 2

n−

1

+

n−

1

X

k=0

1 − ρ ε 2

k

!

χ

{ρ<ε2}

,

(9)

(f

nε

)

′′

(ρ) = − 1

ρ 2 χ

{ρ≥ε2}

− n 2 − 1 ε 4 1 − ρ

ε 2

n−

2

+ 1 ε 4

n

X

3

k=0

(k + 1) 1 − ρ ε 2

k

!

χ

{ρ<ε2}

,

which immediately yields that

| √ ρ(f

nε

)

(ρ) | ≤ 2n ε ,

| ρ 3/2 (f

nε

)

′′

(ρ) | ≤ 1

ε n 2 − 1 +

n

X

3

k=0

(k + 1)

!

= 3n(n − 1) 2ε < 3n 2

2ε . For ρ ∈ [0, ε 2 ], in view of ε ∈ (0, 1], one deduces

| f

nε

(ρ) | ≤ ln(ε

2 ) + n + 1 n

1 − ρ

ε 2

n

+

n

X

1

k=1

1 k

1 − ρ

ε 2

k

≤ ln(ε

2 ) + n + 1

n +

n−

1

X

k=1

1 k

≤ ln(ε

2 ) + 2 + X

n

k=2

1 k

≤ 2 + ln(nε

2 ),

which together with | f

nε

(ρ) | ≤ max { ln(ε

2 ), | ln(A) |} when ρ ∈ [ε 2 , A] concludes

(2.10).

2.3. Comparison between different regularizations. To compare different reg- ularizations for F(ρ) (and thus for f (ρ)), Fig. 1 shows F

nε

(n = 2, 4, 100, 500), F e

ε

and F b

ε

for different ε, from which we can see that the newly proposed local regu- larization F

nε

approximates F more accurately.

Fig. 2 shows various regularizations f

nε

(n = 2, 4, 100, 500), f e

ε

and f b

ε

for various ε, while Figs. 3 & 4 show their first- and second-order derivatives. From these figures, we can see that the newly proposed local regularization f

nε

(and its deriva- tives with larger n) approximates the nonlinearity f (and its derivatives) more accurately. In addition, Fig. 5 depicts F

nε

(ρ) (with ε = 0.1) and its derivatives for different n, from which we can clearly see the convergence of F

nε

(ρ) (and its derivatives) to F(ρ) (and its derivatives) W.R.T. order n.

3. Local energy regularization (LER) for the LogNLS In this section, we consider the regularized energy

(3.1) E

nε

(u) :=

Z

|∇ u | 2 + λF

nε

( | u | 2 ) dx,

where F

nε

is defined in (2.1). The Hamiltonian flow of the regularized energy i∂

t

u =

δEnε

(u)

δu

yields the following energy regularized logarithmic Schr¨odinger equation (ERLogSE) with a regularizing parameter 0 < ε ≪ 1,

(3.2)

( i∂

t

u

ε

(x, t) = − ∆u

ε

(x, t) + λ u

ε

(x, t) f

nε

( | u

ε

(x, t) | 2 ), x ∈ Ω, t > 0, u

ε

(x, 0) = u 0 (x), x ∈ Ω.

We recall that f

nε

is defined by (2.6).

(10)

Figure 1. Comparison of different regularizations for F (ρ) = ρ ln ρ − ρ.

3.1. The Cauchy problem. To investigate the well-posedness of the problem (3.2), we first introduce some appropriate spaces. For α > 0 and Ω = R

d

, denote by L 2

α

the weighted L 2 space

L 2

α

:= { v ∈ L 2 ( R

d

), x 7−→ h x i

α

v( x ) ∈ L 2 ( R

d

) } , where h x i := p

1 + | x | 2 , with norm k v k

L2α

:= kh x i

α

v( x ) k

L2

(

Rd

) . Regarding the Cauchy problem (3.2), we have similar results as for the regularization (1.6) in [5], but not quite the same. For the convenience of the reader, we recall the main arguments.

Theorem 3.1. Let λ ∈ R , u 0 ∈ H 1 (Ω), and 0 < ε ≤ 1.

(1). For (3.2) posed on Ω = R

d

or a bounded domainwith homogeneous Dirichlet

or periodic boundary condition, there exists a unique, global weak solution u

ε

(11)

Figure 2. Comparison of different regularizations for the nonlin- earity f (ρ) = ln ρ.

L

loc ( R ; H 1 (Ω)) to (3.2) (with H 0 1 (Ω) instead of H 1 (Ω) in the Dirichlet case). Fur- thermore, for any given T > 0, there exists a positive constant C(λ, T ) (independent of n) such that

(3.3) k u

ε

k

L

([0,T ];H

1

(Ω)) ≤ C(λ, T ) k u 0 k

H1

(Ω) , ∀ ε > 0.

(2). For (3.2) posed on a bounded domainwith homogeneous Dirichlet or periodic boundary condition, if in addition u 0 ∈ H 2 (Ω), then u

ε

∈ L

loc ( R ; H 2 (Ω)) and there exists a positive constant C(n, λ, T ) such that

(3.4) k u

ε

k

L

([0,T];H

2

(Ω)) ≤ C(n, λ, T ) k u 0 k

H2

(Ω) , ∀ ε > 0.

(3). For (3.2) on Ω = R

d

, suppose moreover u 0 ∈ L 2

α

, for some 0 < α ≤ 1.

(12)

Figure 3. Comparison of different regularizations for f

(ρ) = 1/ρ.

There exists a unique, global weak solution u

ε

∈ L

loc ( R ; H 1 ( R

d

) ∩ L 2

α

) to (3.2), and

k u

ε

k

L

([0,T];H

1

) ≤ C(n, λ, T ) k u 0 k

H1

,

k u

ε

k

L

([0,T];L

2α

) ≤ C(n, λ, T, k u 0 k

H1

) k u 0 k

L2α

, ∀ ε > 0.

(3.5)

If in addition u 0 ∈ H 2 ( R

d

), then u

ε

∈ L

loc ( R ; H 2 ( R

d

)), and (3.6) k u

ε

k

L

([0,T];H

2

) ≤ C(n, λ, T, k u 0 k

H2

, k u 0 k

L2α

), ∀ ε > 0.

If u 0 ∈ H 2 ( R

d

) ∩ L 2 2 , then u

ε

∈ L

loc ( R ; H 2 ( R

d

) ∩ L 2 2 ).

Proof. (1). For fixed ε > 0, the nonlinearity in (3.2) is locally Lipschitz contin-

uous, and grows more slowly than any power of | u

ε

| . Standard Cauchy theory

for nonlinear Schr¨odinger equations implies that there exists a unique solution

u

ε

∈ L

loc ( R ; H 1 (Ω)) to (3.2) (respectively, u

ε

∈ L

loc ( R ; H 0 1 (Ω)) in the Dirichlet

case); see e.g. [19, Corollary 3.3.11 and Theorem 3.4.1]. In addition, the L 2 -norm

(13)

Figure 4. Comparison of different regularizations for f

′′

(ρ) =

− 1/ρ 2 .

of u

ε

is independent of time,

k u

ε

(t) k 2

L2

(Ω) = k u 0 k 2

L2

(Ω) , ∀ t ∈ R . For j ∈ { 1, . . . , d } , differentiate (3.2) with respect to x

j

:

(i∂

t

+ ∆) ∂

j

u

ε

= λ∂

j

u

ε

f

nε

( | u

ε

| 2 ) + 2λu

ε

(f

nε

)

( | u

ε

| 2 )Re (u

ε

j

u

ε

) .

Multiply the above equation by ∂

j

u

ε

, integrate on Ω, and take the imaginary part:

(2.9) implies

1 2

d

dt k ∂

j

u

ε

k 2

L2

(Ω) ≤ 6 | λ |k ∂

j

u

ε

k 2

L2

(Ω) , hence (3.3), by Gronwall lemma.

(2). The propagation of the H 2 regularity is standard, since f

nε

is smooth, so we

focus on (3.4). We now differentiate (3.2) with respect to time: we get the same

(14)

Figure 5. Comparison of regularizations g

n

0.1 (g = F, f, f

, f

′′

) with different order n.

estimate as above, with ∂

j

replaced by ∂

t

, and so

k ∂

t

u

ε

(t) k 2

L2

(Ω) ≤ k ∂

t

u

ε

(0) k 2

L2

(Ω) e 12

|λ||t|

. In view of (3.2),

i∂

t

u

ε|t=0

= − ∆u 0 + λu 0 f

nε

( | u 0 | 2 ).

For 0 < δ < 1, we have

√ ρ | f

nε

(ρ) | ≤ C(δ)

ρ 1/2

δ/2

+ ρ 1/2+δ/2 ,

for some C(δ) independent of ε and n, so for δ > 0 sufficiently small, Sobolev embedding entails

k ∂

t

u

ε

(0) k

L2

(Ω) ≤ k u 0 k

H2

(Ω) + C(δ)

k u 0 k 1

L2−2δδ

(Ω) + k u 0 k 1+δ

H1

(Ω)

. Since Ω is bounded, H¨ older inequality yields

k u 0 k

L2−2δ

(Ω) ≤ k u 0 k

L2

(Ω | Ω |

δ/(2

2δ) .

Thus, the first term in (3.2) is controlled in L 2 . Using the same estimates as above, we control the last term in (3.2) (thanks to (3.3)), and we infer an L 2 -estimate for

∆u

ε

, hence (3.4).

(3). In the case Ω = R

d

, we multiply (3.2) by h x i

α

, and the same energy estimate as before now yields

d

dt k u

ε

k 2

L2α

= 4α Im Z

Rd

x · ∇ u

ε

h x i 2

u

ε

(t) d x . k h x i

1 u

ε

k

L2

(

Rd

) k∇ u

ε

k

L2

(

Rd

)

. k h x i

α

u

ε

k

L2

(R

d

) k∇ u

ε

k

L2

(R

d

) ,

(15)

where the last inequality follows from the assumption α ≤ 1, hence (3.5). To prove (3.6), we resume the same approach as to get (3.4), with the difference that the H¨ older estimate must be replaced by some other estimate (see e.g. [17]): for δ > 0 sufficiently small,

Z

Rd

| u | 2

. k u k 2

L2

(R

d

)

dδ/α

k| x |

α

u k

dδ/αL2

(R

d

) .

The L 2 2 estimate follows easily, see e.g. [5] for details.

3.2. Convergence of the regularized model. In this subsection, we show an approximation property of the regularized model (3.2) to (1.1).

Lemma 3.2. Suppose the equation (3.2) is set on Ω, where Ω = R d , or Ω ⊂ R d is a bounded domain with homogeneous Dirichlet or periodic boundary condition. We have the general estimate:

(3.7) d

dt k u ε (t) − u(t) k 2 L

2

≤ | λ | 4 k u ε (t) − u(t) k 2 L

2

+ 6ε k u ε (t) − u(t) k L

1

. Proof. Subtracting (1.1) from (3.2), we see that the error function e ε := u ε − u satisfies

i∂ t e ε + ∆e ε = λ

u ε ln( | u ε | 2 ) − u ln( | u | 2 ) + λu ε

f n ε ( | u ε | 2 ) − ln( | u ε | 2 )

χ

{|

u

ε|

}

. Multiplying the above error equation by e ε (t), integrating in space and taking imaginary parts, we can get by using Lemma 2.1, (2.4) and (2.5) that

1 2

d

dt k e ε (t) k 2 L

2

= 2λ Im Z

[u ε ln( | u ε | ) − u ln( | u | )] e ε (x, t)dx + λ Im

Z

|

u

ε|

u ε

f n ε ( | u ε | 2 ) − ln( | u ε | 2 )

e ε (x, t)dx

≤ 2 | λ |k e ε (t) k 2 L

2

+ | λ | Z

|

u

ε|

u ε e ε

Q ε n ( | u ε | 2 ) − ln( | u ε | 2 ) + | u ε | 2 (Q ε n )

( | u ε | 2 ) dx

≤ 2 | λ |k e ε (t) k 2 L

2

+ | λ | Z

|

u

ε|

e ε u ε h Z ε

2

|

u

ε|2

(s − | u ε | 2 ) n

s n+1 ds − 1 + | u ε | 2 (Q ε n )

( | u ε | 2 ) i dx

= 2 | λ | k e ε (t) k 2 L

2

+ | λ | Z

|

u

ε|

e ε u ε h Z ε

2

|

u

ε|2

(s − | u ε | 2 ) n s n+1 ds −

1 − | u ε | 2 ε 2

n i dx

≤ 2 | λ | k e ε (t) k 2 L

2

+ ε | λ |k e ε k L

1

+ | λ | Z ε

2

0

s

n

1 Z

|

u

ε|2

<s

e ε u ε (s − | u ε | 2 ) n dxds

≤ 2 | λ | k e ε (t) k 2 L

2

+ 3ε | λ |k e ε k L

1

.

This yields the result.

Invoking the same arguments as in [5], based on the previous error estimate, and interpolation between L 2 and H 2 , we get the following error estimate.

Proposition 3.3. Ifhas finite measure and u 0 ∈ H 2 (Ω), then for any T > 0,

k u ε − u k L

([0,T];L

2

(Ω)) ≤ C 1 ε, k u ε − u k L

([0,T];H

1

(Ω)) ≤ C 2 ε 1/2 ,

(16)

where C 1 depends on | λ | , T , | Ω | , and C 2 depends in addition on k u 0 k H

2

(Ω) . If Ω = R d , 1 ≤ d ≤ 3 and u 0 ∈ H 2 ( R d ) ∩ L 2 2 , then for any T > 0, we have

k u ε − u k L

([0,T];L

2

(

Rd

)) ≤ D 1 ε

4+d4

, k u ε − u k L

([0,T];H

1

(

Rd

)) ≤ D 2 ε

4+d2

, where D 1 and D 2 depend on d, | λ | , T , k u 0 k L

22

and k u 0 k H

2

(

Rd

) .

Proof. The proof is the same as that in [5]. We just list the outline for the readers’

convenience. When Ω is bounded, the convergence in L 2 follows from Gronwall’s inequality by applying (3.7) and the estimate k v k L

1

≤ | Ω | 1/2 k v k L

2

. The estimate in H 1 follows form the Gagliardo-Nirenberg inequality k v k H

1

≤ C k v k 1/2 L

2

k v k 1/2 H

2

and the property (3.4). For Ω = R d , the convergence in L 2 can be established by Gronwall’s inequality and the estimate (cf. [5])

k v k L

1

≤ C d k v k 1 L

2

d/4 k v k d/4 L

22

≤ C d

ε

1 k v k 2 L

2

+ ε

4−d4+d

k v k

2d 4+d

L

22

,

which is derived by the Cauchy-Schwarz inequality and Young’s inequality. The convergence in H 1 can similarly derived by the Gagliardo-Nirenberg inequality.

3.3. Convergence of the energy. By construction, the energy is conserved, i.e., (3.8) E n ε (u ε ) =

Z

|∇ u ε (x, t) | 2 + λF n ε ( | u ε (x, t) | 2 )

dx = E n ε (u 0 ).

For the convergence of the energy, we have the following estimate.

Proposition 3.4. For u 0 ∈ H 1 (Ω) ∩ L α (Ω) with α ∈ (0, 2), the energy E n ε (u 0 ) converges to E(u 0 ) with

| E n ε (u 0 ) − E(u 0 ) | ≤ | λ | k u 0 k α L

α

ε 2

α 1 − α/2 . In addition, for bounded Ω, we have

| E n ε (u 0 ) − E(u 0 ) | ≤ | λ | | Ω | ε 2 .

Proof. It can be deduced from the definition (3.8) and (2.4) that

| E n ε (u 0 ) − E(u 0 ) | = | λ | Z

[F ( | u 0 (x) | 2 ) − F n ε ( | u 0 (x) | 2 )]dx

= | λ | Z

|

u

0

(

x

)

|

<ε | u 0 (x) | 2 [Q( | u 0 (x) | 2 ) − Q ε n ( | u 0 (x) | 2 )]dx

= | λ | Z

|

u

0

(

x

)

|

<ε | u 0 (x) | 2 Z ε

2

|

u

0

(

x

)

|2

s

n

1 (s − | u 0 (x) | 2 ) n dsdx

= | λ | Z ε

2

0

s

n

1 Z

|

u

0

(

x

)

|2

<s | u 0 (x) | 2 (s − | u 0 (x) | 2 ) n dxds.

If Ω is bounded, we immediately get

| E n ε (u 0 ) − E(u 0 ) | ≤ | λ | | Ω | ε 2 . For unbounded Ω, one gets

| E n ε (u 0 ) − E(u 0 ) | ≤ | λ | Z ε

2

0

s

n

1 s n+1

α/2 k u 0 k α L

α

ds = | λ | k u 0 k α L

α

ε 2

α 1 − α/2 ,

which completes the proof.

(17)

Remark 3.5. Recall that it was shown in [5] that for the regularized model (1.6) with the energy density (1.11), the energy

(3.9) E e ε (u 0 ) = k∇ u 0 k 2 L

2

+ λ Z

F e ε ( | u 0 | 2 )dx

converges to E(u 0 ) with an error O(ε). For the regularization (1.7) with the energy density (1.11) and the regularized energy

(3.10) E b ε (u 0 ) = k∇ u 0 k 2 L

2

+ λ Z

F b ε ( | u 0 | 2 )dx, we have

b E ε (u 0 ) − E(u 0 ) = | λ | Z

[F( | u 0 (x) | 2 ) − F b ε ( | u 0 (x) | 2 )]dx

= | λ | Z

2 + | u 0 | 2 ) ln(ε 2 + | u 0 | 2 ) − ε 2 ln(ε 2 ) − | u 0 | 2 ln( | u 0 | 2 ) dx

≤ | λ | ε 2 Z

ln

1 + | u 0 | 2 ε 2

d x + | λ | Z

Ω | u 0 | 2 ln

1 + ε 2

| u 0 | 2

d x

≤ | λ | ε 2

α C(α) Z

Ω | u 0 | α dx

= | λ | ε 2

α C(α) k u 0 k α L

α

,

where we have used the inequality ln(1 + x) ≤ C(β)x β for β ∈ (0, 1] and x ≥ 0.

Hence for u 0 ∈ H 1 (Ω) ∩ L α (Ω) with α ∈ (0, 2), we infer

b E ε (u 0 ) − E(u 0 ) ≤ | λ | ε 2

α C(α) k u 0 k α L

α

,

that is, the same convergence rate as E n ε . Thus the newly proposed local energy regularization F n ε is more accurate than F e ε , and than F b ε in the case of bounded domains, from the viewpoint of energy.

4. Regularized Lie-Trotter splitting methods

In this section, we investigate approximation properties of the Lie-Trotter split- ting methods [13, 27, 47] for solving the regularized model (3.2) in one dimension (1D). Extensions to higher dimensions are straightforward. To simplify notations, we set λ = 1.

4.1. A time-splitting for (3.2). The operator splitting methods are based on a decomposition of the flow of (3.2):

∂ t u ε = A(u ε ) + B(u ε ), where

A(v) = i∆v, B(v) = − ivf n ε ( | v | 2 ), and the solution of the sub-equations

(4.1)

( ∂ t v(x, t) = A(v(x, t)), x ∈ Ω, t > 0, v(x, 0) = v 0 (x),

(4.2)

( ∂ t ω(x, t) = B(ω(x, t)), x ∈ Ω, t > 0,

ω(x, 0) = ω 0 (x),

(18)

where Ω = R or Ω ⊂ R is a bounded domain with homogeneous Dirichlet or periodic boundary condition on the boundary. Denote the flow of (4.1) and (4.2) by (4.3) v( · , t) = Φ t A (v 0 ) = e it∆ v 0 , ω( · , t) = Φ t B (ω 0 ) = ω 0 e

itf

nε

(

|

ω

0|2

) , t ≥ 0.

As is well known, the flow Φ t A satisfies the isometry relation (4.4) k Φ t A (v 0 ) k H

s

= k v 0 k H

s

, ∀ s ∈ R , ∀ t ≥ 0.

Regarding the flow Φ t B , we have the following properties.

Lemma 4.1. Assume τ > 0 and ω 0 ∈ H 1 (Ω), then

(4.5) k Φ τ B (ω 0 ) k L

2

= k ω 0 k L

2

, k Φ τ B (ω 0 ) k H

1

≤ (1 + 6τ) k ω 0 k H

1

. For v, w ∈ L 2 (Ω),

(4.6) k Φ τ B (v) − Φ τ B (w) k L

2

≤ (1 + 4nτ) k v − w k L

2

. Proof. By direct calculation, we get

∂ x Φ τ B (ω 0 ) = e

iτ f

nε

(

|

ω

0|

)

2

∂ x ω 0 − iτ (f n ε )

( | ω 0 | 2 )(ω 2 0 ∂ x ω 0 + | ω 0 | 2 ∂ x ω 0 ) , which immediately gives (4.5) by recalling (2.9). We claim that for any x ∈ Ω,

| Φ τ B (v)(x) − Φ τ B (w)(x) | ≤ (1 + 4nτ) | v(x) − w(x) | .

Assuming, for example, | v(x) | ≤ | w(x) | , by inserting a term v(x)e

iτ f

nε

(

|

w(x)

|

)

2

, we can get

| Φ τ B (v)(x) − Φ τ B (w)(x) |

= v(x)e

iτ f

nε

(

|

v(x)

|

)

2

− w(x)e

iτ f

nε

(

|

w(x)

|

)

2

= v(x) − w(x) + v(x)

e iτ[f

nε

(

|

w(x)

|2

)

f

nε

(

|

v(x)

|2

)] − 1

≤ | v(x) − w(x) | + 2 | v(x) | sin τ 2

f n ε ( | w(x) | 2 ) − f n ε ( | v(x) | 2 )

≤ | v(x) − w(x) | + τ | v(x) | | f n ε ( | w(x) | 2 ) − f n ε ( | v(x) | 2 ) |

≤ (1 + 4nτ) | v(x) − w(x) | ,

where we have used the estimate (2.7). When | v(x) | ≥ | w(x) | , the same inequality can be obtained by exchanging v and w in the above computation. Thus the proof

for (4.6) is complete.

4.2. Error estimates for Φ τ = Φ τ A Φ τ B . We consider the Lie-Trotter splitting (4.7) u ε,k+1 = Φ τ (u ε,k ) = Φ τ Aτ B (u ε,k )), k ≥ 0; u ε,0 = u 0 , τ > 0.

For u 0 ∈ H 1 (Ω), it follows from (4.4) and (4.5) that k u ε,k k L

2

= k u ε,k

1 k L

2

≡ k u ε,0 k L

2

= k u 0 k L

2

,

k u ε,k k H

1

≤ (1 + 6τ) k u ε,k

1 k H

1

≤ e 6kτ k u 0 k H

1

, k ≥ 0.

(4.8)

Theorem 4.2. Let T > 0 and τ 0 > 0 be given constants. Assume that the solution of (3.2) satisfies u ε ∈ L

([0, T ]; H 1 (Ω)) and the time step τ ≤ τ 0 . Then there exists 0 < ε 0 < 1 depending on n, τ 0 and M := k u ε k L

([0,T ];H

1

(Ω)) such that when ε ≤ ε 0 and t k := kτ ≤ T , we have

(4.9) k u ε,k − u ε (t k ) k L

2

≤ C (n, τ 0 , T, M) ln(ε

11/2 .

(19)

Proof. Denote the exact flow of (3.2) by u ε (t) = Ψ t (u 0 ). First, we establish the local error for v ∈ H 1 (Ω):

(4.10) k Ψ τ (v) − Φ τ (v) k L

2

≤ C(n, τ 0 ) k v k H

1

ln(ε

13/2 , τ ≤ τ 0 , when ε is sufficiently small. Note that definitions imply

i∂ t Ψ t (v) + ∆Ψ t (v) = Ψ t (v)f n ε ( | Ψ t (v) | 2 ), i∂ t Φ t (v) + ∆Φ t (v) = Φ t A Φ t B (v)f n ε ( | Φ t B (v) | 2 )

. Denoting E t (v) = Ψ t (v) − Φ t (v), we have

(4.11) i∂ t E t (v) + ∆ E t (v) = Ψ t (v)f n ε ( | Ψ t (v) | 2 ) − Φ t A Φ t B (v)f n ε ( | Φ t B (v) | 2 ) . Multiplying (4.11) by E t (v), integrating in space and taking the imaginary part, we get

1 2

d

dt kE t (v) k 2 L

2

= Im Ψ t (v)f n ε ( | Ψ t (v) | 2 ) − Φ t A Φ t B (v)f n ε ( | Φ t B (v) | 2 ) , E t (v)

= Im Ψ t (v)f n ε ( | Ψ t (v) | 2 ) − Φ t (v)f n ε ( | Φ t (v) | 2 ), E t (v) + Im Φ t (v)f n ε ( | Φ t (v) | 2 ) − Φ t A Φ t B (v)f n ε ( | Φ t B (v) | 2 )

, E t (v)

≤ 4n kE t (v) k 2 L

2

+ Φ t (v)f n ε ( | Φ t (v) | 2 ) − Φ t A Φ t B (v)f n ε ( | Φ t B (v) | 2 ) L

2

kE t (v) k L

2

, where we have used (2.8) and the scalar product is the standard one in L 2 : (u, w) = R

Ω u(x)w(x)dx. This implies

(4.12) d

dt kE t (v) k L

2

≤ 4n kE t (v) k L

2

+ J 1 + J 2 , where

J 1 = k Φ t (v)f n ε ( | Φ t (v) | 2 ) − Φ t B (v)f n ε ( | Φ t B (v) | 2 ) k L

2

, J 2 = k Φ t B (v)f n ε ( | Φ t B (v) | 2 ) − Φ t A Φ t B (v)f n ε ( | Φ t B (v) | 2 )

k L

2

.

To estimate J 1 in (4.12), first we try to find the bound of k Φ t (v) k L

, k Φ t B (v) k L

. It follows from (4.4) and (4.5) that

k Φ t (v) k H

1

= k Φ t B (v) k H

1

≤ (1 + 6t) k v k H

1

≤ (1 + 6t 0 ) k v k H

1

, t ≤ t 0 . Hence by Sobolev embedding, we have

(4.13) k Φ t (v) k L

≤ c(1 + 6t 0 ) k v k H

1

, k Φ t B (v) k L

≤ c(1 + 6t 0 ) k v k H

1

, where c is the constant in the Sobolev inequality k ω k L

≤ c k ω k H

1

. Next we claim that for y, z satisfying | y | , | z | ≤ D, it can be established that

(4.14) | yf n ε ( | y | 2 ) − zf n ε ( | z | 2 ) | ≤ 4 ln(ε

1 ) | y − z | ,

when ε is sufficiently small. It follows from (2.10) that | f n ε ( | y | 2 ) | ≤ 2 + ln(nε

2 ), when | y | ≤ D and ε ≤ √

n/D. Assuming, for example, 0 < | z | ≤ | y | , and applying (2.7), we get

| yf n ε ( | y | 2 ) − zf n ε ( | z | 2 ) | = | (y − z)f n ε ( | y | 2 ) | + | z || f n ε ( | y | 2 ) − f n ε ( | z | 2 ) |

≤ (2 + ln(nε

2 )) | y − z | + | z | 4n | y − z |

| z |

≤ 2(3n + ln(ε

1 )) | y − z |

(20)

≤ 4 ln(ε

1 ) | y − z | , when ε ≤ ε e := min { √

n/D, e

3n } . The case when y = 0 or z = 0 can be handled similarly. Recalling (4.13), taking D = c(1 + 6t 0 ) k v k H

1

, we obtain, when ε ≤ ε 1 :=

min { c(1+6t

0

n )

k

v

kH1

, e

3n } ,

J 1 ≤ 4 ln(ε

1 ) k Φ t (v) − Φ t B (v) k L

2

≤ 4 ln(ε

1 ) √

2t k Φ t B (v) k H

1

≤ 6 ln(ε

1 ) √ t k v k H

1

, (4.15)

where we have used the estimate

(4.16) ω − Φ t A (ω)

L

2

≤ √

2t k ω k H

1

, as in [4], instead of the estimate from [13],

ω − Φ t A (ω)

L

2

≤ 2t k ω k H

2

,

which in our case yields an extra 1/ε factor in the error estimate.

To estimate J 2 , we first claim that

(4.17) k Φ t B (v)f n ε ( | Φ t B (v) | 2 ) k H

1

≤ 6 ln(ε

1 )(1 + 3t 0 ) k v k H

1

, when ε ≤ ε 1 and t ≤ t 0 . Recalling that

Φ t B (v)f n ε ( | Φ t B (v) | 2 ) = vf n ε ( | v | 2 )e

itf

nε

(

|

v

|2

) , and | f n ε ( | v | 2 ) | ≤ 3 ln(ε

1 ), when ε ≤ ε 1 , this implies

k (Φ t B (v))f n ε ( | Φ t B (v) | 2 ) k L

2

≤ 3 ln(ε

1 ) k v || L

2

. Noticing that

∂ x [Φ t B (v)f n ε ( | Φ t B (v) | 2 )] = e

itf

nε

(

|

v

|2

)

v x f n ε ( | v | 2 )

+(1 − itf n ε ( | v | 2 ))(f n ε )

( | v | 2 )(v 2 v x + | v | 2 v x ) , which together with (2.9) yields

| ∂ x [Φ t B (v)f n ε ( | Φ t B (v) | 2 )] | ≤

6 + 3 ln(ε

1 )(1 + 6t 0 )

| v x | ≤ 6 ln(ε

1 )(1 + 3t 0 ) | v x | , which immediately gives (4.17). Applying (4.16) again entails

(4.18) J 2 ≤ √

2t k (Φ t B (v))f n ε ( | Φ t B (v) | 2 ) k H

1

≤ 9 ln(ε

1 )(1 + 3t 0 ) √ t k v k H

1

, for ε ≤ ε 1 and t ≤ t 0 . Combining (4.12), (4.15) and (4.18), we get

d

dt kE t (v) k L

2

≤ 4n kE t (v) k L

2

+ 15(1 + 2t 0 ) ln(ε

1 )) √ t k v k H

1

. Invoking Gronwall’s inequality, we have

kE τ (v) k L

2

≤ e 4nτ

kE 0 (v) k L

2

+ 15(1 + 2τ 0 ) ln(ε

1 ) k v k H

1

Z τ 0

√ sds

≤ 30(1 + 2τ 0 )e 4nτ k v k H

1

ln(ε

13/2

≤ C(n, τ 0 ) k v k H

1

ln(ε

13/2 ,

when τ ≤ τ 0 and ε ≤ ε 0 := min { c(1+6τ

n

0

)M , e

3n } depending on τ 0 , n and M = k u ε k L

([0,T];H

1

) , which completes the proof for (4.10).

Next we infer the stability analysis for the operator Φ t :

(4.19) k Φ τ (v) − Φ τ (w) k L

2

≤ (1 + 4nτ) k v − w k L

2

, for v, w ∈ L 2 (Ω).

Références

Documents relatifs

In order to suppress the round-off error and to avoid blow-up, a regularized logarithmic Schr¨ odinger equation (RLogSE) is proposed with a small regularization parameter 0 &lt; ε ≪

Remaud , General properties of Gausson-conserving descriptions of quantal damped motion, Physica A, 105 (1980), pp. Makridakis , A space-time finite element method for the

This paper is devoted to the proof of the convergence of the solution of the Boltzmann equation, for a degenerate semiconductor and with an arbitrary band structure, towards

, On smoothing property of Schr¨ odinger propagators, in Functional- Analytic Methods for Partial Differential Equations (Tokyo, 1989 ), Lecture Notes in Math.. , Smoothing property

Logarithmic Choquard equation; planar Schrödinger–Newton system; positive ground state solution; asymptotics;

Estimates in dimension d = 2 for constant magnetic fields Numerical results and the symmetry issue. 3 Magnetic rings: the one-dimensional periodic case Magnetic interpolation on

As Duyckaerts [6] did for the simplified model, the idea is, first, to use a known result of Burq (see [5]) which links, for dissipative operators, logarithmic decay to

Pour contrˆ oler le terme nonlin´ eaire, on utilise les deux outils. In´ egalit´ es