In this article, we obtain new results on the unconditional well-posedness for a pair of periodic nonlinear dispersive equations using an abstract framework introduced by Kishimoto. This framework is based on a normal form reductions argument coupled with a number of crucial multilinear estimates.
Citation: Dan-Andrei Geba. Unconditional well-posedness for the periodic Boussinesq and Kawahara equations[J]. Electronic Research Archive, 2024, 32(2): 1067-1081. doi: 10.3934/era.2024052
[1] | Dan-Andrei Geba, Evan Witz . Revisited bilinear Schrödinger estimates with applications to generalized Boussinesq equations. Electronic Research Archive, 2020, 28(2): 627-649. doi: 10.3934/era.2020033 |
[2] | Xiuli Xu, Lian Yang . Global well-posedness of the 3D nonlinearly damped Boussinesq magneto-micropolar system without heat diffusion. Electronic Research Archive, 2025, 33(4): 2285-2294. doi: 10.3934/era.2025100 |
[3] | Hua Qiu, Zheng-An Yao . The regularized Boussinesq equations with partial dissipations in dimension two. Electronic Research Archive, 2020, 28(4): 1375-1393. doi: 10.3934/era.2020073 |
[4] | Lin Shen, Shu Wang, Yongxin Wang . The well-posedness and regularity of a rotating blades equation. Electronic Research Archive, 2020, 28(2): 691-719. doi: 10.3934/era.2020036 |
[5] | Lianbing She, Nan Liu, Xin Li, Renhai Wang . Three types of weak pullback attractors for lattice pseudo-parabolic equations driven by locally Lipschitz noise. Electronic Research Archive, 2021, 29(5): 3097-3119. doi: 10.3934/era.2021028 |
[6] | Liju Yu, Jingjun Zhang . Global solution to the complex short pulse equation. Electronic Research Archive, 2024, 32(8): 4809-4827. doi: 10.3934/era.2024220 |
[7] | Nisachon Kumankat, Kanognudge Wuttanachamsri . Well-posedness of generalized Stokes-Brinkman equations modeling moving solid phases. Electronic Research Archive, 2023, 31(3): 1641-1661. doi: 10.3934/era.2023085 |
[8] | Changjia Wang, Yuxi Duan . Well-posedness for heat conducting non-Newtonian micropolar fluid equations. Electronic Research Archive, 2024, 32(2): 897-914. doi: 10.3934/era.2024043 |
[9] | Yong Zhou, Jia Wei He, Ahmed Alsaedi, Bashir Ahmad . The well-posedness for semilinear time fractional wave equations on $ \mathbb R^N $. Electronic Research Archive, 2022, 30(8): 2981-3003. doi: 10.3934/era.2022151 |
[10] | Junseok Kim . Maximum principle preserving the unconditionally stable method for the Allen–Cahn equation with a high-order potential. Electronic Research Archive, 2025, 33(1): 433-446. doi: 10.3934/era.2025021 |
In this article, we obtain new results on the unconditional well-posedness for a pair of periodic nonlinear dispersive equations using an abstract framework introduced by Kishimoto. This framework is based on a normal form reductions argument coupled with a number of crucial multilinear estimates.
In this paper, we study the periodic Cauchy problems
{∂2tu−∂2xu+∂4xu+∂2x(u2)=0,u=u(t,x):R×T→R,u(0,x)=u0(x), ∂tu(0,x)=u1(x),(u0,u1)∈Hs(T)×Hs−2(T), | (1.1) |
and
{∂tu+β∂3xu−∂5xu+∂x(u2)=0,u=u(t,x):R×T→R,u(0,x)=u0(x),u0∈Hs(T), | (1.2) |
where
u(t,0)=u(t,2π),T=R/2πZ,andβ=0,1,or−1. |
The first partial differential equation is called the "good" Boussinesq equation and it is known to describe electromagnetic waves in nonlinear dielectrics [1]. When the quadratic nonlinearity is replaced by 4u3−6u5, the resulting equation was, in fact, derived in the context of shape-memory alloys [2]. The "good" Boussinesq equation can be seen as an improved frequency dispersion version of the original water wave equation derived by Boussinesq [3],
∂2tu−gh∂2xu−gh∂2x{32hu2+h23∂2xu}=0, |
where g is the gravitational acceleration, h is undisturbed depth of the channel where the water waves are observed, and u=u(t,x) is the elevation of the water surface. This equation admits solitary wave solutions of the form
u(t,x)=asech2[√3ah3(x±ct)], |
where a and c are suitably chosen constants, and many consider this to be the place where the stability theory of solitary wave solutions started. For a detailed review of various mathematical and physical aspects concerning Boussinesq equations and their generalizations, we refer the reader to the excellent review article by Makhankov [4].
The partial differential equation in the second Cauchy problem is referred to as the Kawahara equation and it represents a scalar approximation for the full water wave equations in the shallow water regime, when one also takes into account the surface tension. For the derivation of this type of equations, one uses three non-dimensional parameters: δ and ϵ which quantize the dispersive and nonlinear effects, and μ, also known as the Bond number. Shallow water regime corresponds to δ≪1 and, if ϵ=δ2 and μ≠1/3, then one obtains
±∂tu+(16−μ2)∂3xu+34∂x(u2)=0, |
which is the well-known KdV equation. On the other hand, if ϵ=δ4 and μ=1/3+νϵ1/2, then one arrives at the Kawahara equation
±∂tu−ν2∂3xu+190∂5xu+34∂x(u2)=0. |
This equation was first derived by Kakutani and Ono [5] to describe nonlinear hydromagnetic waves in plasmas and by Hasimoto [6] precisely in the context of shallow water waves with surface tension and Bond number close to 1/3. For an extensive mathematical and physical perspective on the Kawahara equation, we ask the interested reader to consult the impressive monograph [7] by Lannes.
The well-posedness (WP) of (1.1) and (1.2) has received considerable interest, with the Boussinesq problem being the subject of works by Fang and Grillakis [8], Farah and Scialom [9], Oh and Stefanov [10], Kishimoto [11], Geba et al. [12], and Okamoto [13]. For the Kawahara problem, important contributions were made by Hirayama [14], Kato [15], and Okamoto [13]. The articles by Kishimoto [11] and Kato [15] are of particular significance. The former proves that (1.1) is locally WP for s≥−1/2 and is ill-posed for s<−1/2, while the latter shows that (1.2) is locally WP for s≥−3/2 and is ill-posed for s<−3/2. A parallel, more comprehensive literature addresses the corresponding non-periodic Cauchy problems.
When studying the WP of (1.1) and (1.2), the natural solution spaces, also called Hadamard spaces, are
X=C(I;Hs(T))∩C1(I;Hs−2(T)) |
and
X=C(I;Hs(T)) |
respectively, where I⊆R is a time interval with 0∈I. Yet, in the previously listed WP references, uniqueness of solutions to either (1.1) or (1.2) is only attained in a proper subset of X. This has the format X∩Y, where Y is an additional functional space. It is then natural to ask under what conditions these solutions become unique in their full Hadamard spaces. This is what is called the studying of unconditional uniqueness or, by extension, unconditional well-posedness (UWP) for these Cauchy problems. In this direction, our article provides the following results:
Theorem 1.1. UWP of solutions to the Cauchy problem (1.1) holds for data in Hs(T)×Hs−2(T) when s>0.
Theorem 1.2. UWP of solutions to the Cauchy problem (1.2) holds for data in Hs(T) when s>1/4.
The subject of UWP for nonlinear dispersive equations has produced an impressive amount of research for the last 30 years, arguably starting with Kato's work [16] on nonlinear Schrödinger equations. In the same context, we mention the recent seminal papers of Kwon et al. [17] and Kishimoto [18] addressing fairly generic dispersive problems. Specifically for periodic problems, a selection of notable results consists of the ones obtained by Babin et al. [19], Kwon and Oh [20], Guo et al. [21], Kishimoto [22], and Kato and Tsugawa [23]. For comprehensive discussions on UWP, as well as access to extensive bibliographies on the subject, we refer the reader to [17,18].
Related to our results, we first recall that UWP in the non-periodic case was previously obtained by Farah [24] for the "good" Boussinesq equation and by Geba and Lin [25] for the Kawahara equation, both for data in L2(R). To our knowledge, prior to this paper, there are no UWP arguments in the literature for either (1.1) or (1.2). For both of the sharp locally WP results mentioned before (i.e., Kishimoto [11] for (1.1) and Kato [15] for (1.2)), uniqueness of solutions is derived only in proper subsets of their corresponding Hadamard space. Moreover, it is important to note that the question of what the optimal Sobolev index needs to be in order to guarantee UWP is a fairly delicate issue, as discussed in [17,18]. Current methodologies and previous results on related dispersive equations suggest that s=0 should be the right value for both (1.1) and (1.2).
In proving Theorems 1.1 and 1.2, we apply an abstract framework developed by Kishimoto [18] in the context of nonlinear dispersive equations. This has been successfully used in obtaining UWP for periodic Cauchy problems associated to nonlinear Schrödinger equations [26], the Benjamin-Ono equation [22], and the modified Benjamin-Ono equation [27]. At the heart of this method lies a critical set of multilinear bounds, which are used to derive even more complex multilinear estimates. The latter represent the key elements in an infinite iteration scheme of normal form reductions proving UWP for the Cauchy problem under consideration.
In concluding this section, we want to emphasize that one of the goals of this paper is to advocate for the simplicity and the robustness of Kishimoto's method. These two features allowed us to keep the argument quite concise and, in our opinion, very transparent. Of course, one could argue that, at least for (1.2), our result is likely not optimal, since the Cauchy problem for the KdV equation (which enjoys weaker dispersion when compared to the Kawahara equation) is UWP in L2(T), as proven in [19] through finitely many normal form reductions. Again, our aim is to present a streamlined approach to UWP questions, which caters to a large audience.
First, we write A≲B to stand for A≤CB, where C>0 is a constant varying from line to line and depending on various fixed parameters. Next, we ask that A∼B denotes that both A≲B and B≲A are valid. We also let A≪B signify that A≤ϵB for some small absolute constant ϵ>0.
Secondly, for a function f=f(x) defined on T, its Fourier series is given by
f(x)=∑k∈Zfkeikx,fk=12π∫Te−ikxf(x)dx,∀k∈Z. | (2.1) |
In similar fashion, for v=v(t,x) defined on R×T, we write
v(t,x)=∑k∈Zvk(t)eikx,vk(t)=12π∫Te−ikxv(t,x)dx,∀k∈Z. | (2.2) |
Finally, for s∈R, we will operate with the space
l2s=l2s(Z)={ω=(ωk)k;(⟨k⟩sωk)k∈l2(Z)},‖ω‖2l2s=∑k∈Z⟨k⟩2s|ωk|2, |
where ⟨k⟩=(1+|k|2)1/2.
Here, we present a version of Kishimoto's framework specifically suited to be applied to our two Cauchy problems, (1.1) and (1.2).
In the case of (1.2), we consider the generic Cauchy problem
{∂twk(t)=∑k1+k2=keitφ(k,k1,k2)m(k,k1,k2)wk1(t)wk2(t),w(0)=w(0)∈l2s, | (2.3) |
and we work with the following version of Theorem 1.1 in [18]:
Theorem 2.1. The Cauchy problem (2.3) has at most one solution
w∈C([0,T];l2s) |
if, for some 0<δ<1/2,
‖∑k1+k2=k|m|⟨φ⟩1/2yk1zk2‖l2s≲‖y‖l2s‖z‖l2s, | (2.4) |
‖∑k1+k2=k|m|⟨φ⟩1−δyk1zk2‖l2s−2≲min{‖y‖l2s−2‖z‖l2s,‖y‖l2s‖z‖l2s−2}, | (2.5) |
‖∑k1+k2=k|m|yk1zk2‖l2s−2≲‖y‖l2s‖z‖l2s | (2.6) |
hold true.
For (1.1), we look at the generic Cauchy problem
{∂tvk(t)=∑k1+k2=keitφ11(k,k1,k2)m11(k,k1,k2)vk1(t)vk2(t)+∑k1+k2=keitφ12(k,k1,k2)m12(k,k1,k2)vk1(t)wk2(t)+∑k1+k2=keitφ13(k,k1,k2)m13(k,k1,k2)wk1(t)vk2(t)+∑k1+k2=keitφ14(k,k1,k2)m14(k,k1,k2)wk1(t)wk2(t)+R1[v,w]k,∂twk(t)=∑k1+k2=keitφ21(k,k1,k2)m21(k,k1,k2)vk1(t)vk2(t)+∑k1+k2=keitφ22(k,k1,k2)m22(k,k1,k2)vk1(t)wk2(t)+∑k1+k2=keitφ23(k,k1,k2)m23(k,k1,k2)wk1(t)vk2(t)+∑k1+k2=keitφ24(k,k1,k2)m24(k,k1,k2)wk1(t)wk2(t)+R2[v,w]k,(v(0),w(0))=(v0,w0)∈l2s×l2s. | (2.7) |
In Section 2.5 of [18], Kishimoto discusses how his scheme can be adapted to abstract systems like the one above, and we take as a working version the following result:
Theorem 2.2. The Cauchy problem (2.7) has at most one solution
(v,w)∈C([0,T];l2s×l2s) |
if
‖R1[y,z]‖C([0,T];l2s)+‖R2[y,z]‖C([0,T];l2s)≤C(‖y,z‖C([0,T];l2s)), | (2.8) |
‖R1[y,z]−R1[˜y,˜z]‖C([0,T];l2s)+‖R2[y,z]−R2[˜y,˜z]‖C([0,T];l2s)≤C(‖y,z,˜y,˜z‖C([0,T];l2s))(‖y−˜y‖C([0,T];l2s)+‖z−˜z‖C([0,T];l2s)), | (2.9) |
and, for some 0<δ<1/2,
‖∑k1+k2=k|mij|⟨φij⟩1/2yk1zk2‖l2s≲‖y‖l2s‖z‖l2s, | (2.10) |
‖∑k1+k2=k|mij|⟨φij⟩1−δyk1zk2‖l∞≲min{‖y‖l∞‖z‖l2s,‖y‖l2s‖z‖l∞}, | (2.11) |
‖∑k1+k2=k|mij|yk1zk2‖l∞≲‖y‖l2s‖z‖l2s | (2.12) |
hold true for all 1≤i≤2 and 1≤j≤4.
We begin this section by reformulating the Boussinesq Cauchy problem (1.1) in such a way that we can implement the methodology described before. The first step consists in rewriting it as a Cauchy problem for a Schrödinger equation. As in Kishimoto and Tsugawa [28], if we take
(v,v0):=(u−i(1−∂2x)−1ut,u0−i(1−∂2x)−1u1), |
then (1.1) gets transformed into
{i∂tv−∂2xv=¯v−v2+ω(∂x)(v+¯v2)2,v=v(t,x):R×T→C,v(0,x)=v0(x), | (3.1) |
where
ω(∂x):=−∂2x(1−∂2x)−1. |
Conversely, if v and v0 satisfy this Cauchy problem, then, by letting
(u,u0,u1)=(v+¯v2,v0+¯v02,(1−∂2x)(¯v0−v02i)), |
it is easy to check that u, u0, and u1 are all real-valued and solve (1.1). It is equally important to notice that, for an arbitrary T>0,
U=(C([0,T],Hs)∩C1([0,T],Hs−2))×Hs×Hs−2,V=C([0,T];Hs)×Hs, |
the maps
(u,u0,u1)∈U↦(v,v0)∈V |
and
(v,v0)∈V↦(u,u0,u1)∈U |
are both Lipschitz continuous. Thus, UWP for (1.1) with data in Hs×Hs−2 becomes equivalent to UWP for (3.1) in Hs.
Next, by introducing the Fourier series coefficients for v0 and v according to (2.1) and (2.2), respectively, it follows that (3.1) can be turned into the infinite coupled system of ordinary differential equations
{i∂tvk+k2vk=12(¯v−k−vk)+ω(k)4∑k1,k2∈Zk1+k2=k(vk1+¯v−k1)(vk2+¯v−k2),vk(0)=v0k, |
where
ω(k):=k21+k2. | (3.2) |
If we take
u+k(t):=e−itk2vk(t),u−k(t):=eitk2¯v−k(t), |
then the previous system comes to be
{∂tu+k=i2(u+k−e−2itk2u−k)−iω(k)4∑k1,k2∈Zk1+k2=ke−itk2(eitk21u+k1+e−itk21u−k1)(eitk22u+k2+e−itk22u−k2),∂tu−k=i2(e2itk2u+k−u−k)+iω(k)4∑k1,k2∈Zk1+k2=keitk2(eitk21u+k1+e−itk21u−k1)(eitk22u+k2+e−itk22u−k2),(u+k(0),u−k(0))=(v0k,¯v0,−k). |
Finally, we rely on the notation
{ϕ++(k,k1,k2):=−k2+k21+k22,ϕ+−(k,k1,k2):=−k2+k21−k22,ϕ−+(k,k1,k2):=−k2−k21+k22,ϕ–(k,k1,k2):=−k2−k21−k22, | (3.3) |
and
{R+[y,z]k:=i2(yk−e−2itk2zk),R−[y,z]k:=i2(e2itk2yk−zk), | (3.4) |
to arrive at the following working version of the original Cauchy problem (1.1):
{∂tu±k=∓iω(k)4∑k1,k2∈Zk1+k2=k(e±itϕ++u+k1u+k2+e±itϕ+−u+k1u−k2+e±itϕ−+u−k1u+k2+e±itϕ–u−k1u−k2)+R±[u+,u−]k,(u+k(0),u−k(0))=(v0k,¯v0,−k). | (3.5) |
It is for this system that we verify the validity of Theorem 2.2 when s>0, thus proving Theorem 1.1.
Proposition 3.1. Theorem 2.2 holds true for the Cauchy problem (3.5) when s>0.
Proof. We start the argument by recognizing that, in our setting,
φij∈{ϕ++,ϕ+−,ϕ−+,ϕ–}, mij∈{iω(k)4,−iω(k)4}, {R1,R2}={R+,R−}. |
We deduce directly from (3.4) that
|R+[y,z]k|+|R−[y,z]k|≤|yk|+|zk|,|R+[y,z]k−R+[˜y,˜z]k|+|R−[y,z]k−R−[˜y,˜z]k|≤|yk−˜yk|+|zk−˜zk|, |
which immediately implies (2.8) and (2.9).
From (3.2), we derive
(mij=0⟺k=0)and(|mij|<14). | (3.6) |
Using this fact, the Cauchy-Schwarz inequality, and s≥0, we can easily prove (2.12) as follows:
‖∑k1+k2=k|mij|yk1zk2‖l∞≲‖∑k1+k2=k|yk1||zk2|‖l∞≤‖y‖l2‖z‖l2≤‖y‖l2s‖z‖l2s. |
Next, we turn to the argument for (2.10), which can be reduced, with the help of the Cauchy-Schwarz inequality, to proving
supk(∑k1+k2=k|mij|2⟨φij⟩⟨k⟩2s⟨k1⟩2s⟨k2⟩2s)≲1. |
By relying on (3.6), we can further simplify the previous estimate down to
supk≠0(∑k1+k2=k1⟨φij⟩⟨k⟩2s⟨k1⟩2s⟨k2⟩2s)≲1. | (3.7) |
Now, we introduce the notation
|kmax|=max{|k1|,|k2|},|kmin|=min{|k1|,|k2|}, | (3.8) |
and, based on the triangle inequality, we have
|k|=|k1+k2|≲|kmax|. | (3.9) |
Jointly with (3.3), this bound yields that the size of |φij| is similar to one of
{|kmax||kmin|,|k||kmax|,|k||kmin|,|kmax|2}. |
If kmin=0, then kmax=k and (3.7) follows at once. If kmin≠0, then we split the analysis into two complementary scenarios:
1≤|k|≲|kmax|∼|kmin|,1≤|kmin|≪|k|∼|kmax|. | (3.10) |
If 1≤|k|≲|kmax|∼|kmin|, then the previous fact about the size of |φij| and s>0 lead to
supk≠0(∑k1+k2=k1⟨φij⟩⟨k⟩2s⟨k1⟩2s⟨k2⟩2s)≲supk≠0(∑|k|≲|kmax|1|k||kmax||k|2s|kmax|4s)≲supk≠0(1|k|∑|k|≲|kmax|1|kmax|1+2s)≲1. |
If we are in the second scenario, then, using again the information about the size of |φij| and s≥0, we obtain
supk≠0(∑k1+k2=k1⟨φij⟩⟨k⟩2s⟨k1⟩2s⟨k2⟩2s)≲supk≠0(∑|kmin|≪|k|1|k||kmin|1|kmin|2s)≲supk≠0(∑|kmin|≪|k|1|kmin|2)≲1. |
This completes the proof of (3.7) and, hence, of (2.10).
Finally, we come to the argument for (2.11). First, we use (3.6), the Cauchy-Schwarz inequality, and s≥0 to infer
‖∑k1+k2=k|mij|⟨φij⟩1−δyk1zk2‖l∞≲‖∑k1+k2=k≠01⟨φij⟩1−δ|zk2|‖l∞‖y‖l∞≲supk≠0(∑k1+k2=k1⟨φij⟩2−2δ)1/2‖y‖l∞‖z‖l2≲supk≠0(∑k1+k2=k1⟨φij⟩2−2δ)1/2‖y‖l∞‖z‖l2s. |
This effectively reduces the proof of (2.11) to the one for
supk≠0(∑k1+k2=k1⟨φij⟩2−2δ)≲1. |
Here, we reason in identical fashion with the argument for (3.7), involving the notation (3.8), the information about the size of |φij|, and the complementary scenarios (3.10). The case kmin=0 can be easily dispensed with. Otherwise, if 1≤|k|≲|kmax|∼|kmin|, then
supk≠0(∑k1+k2=k1⟨φij⟩2−2δ)≲supk≠0(1|k|2−2δ∑|k|≲|kmax|1|kmax|2−2δ)≲1, |
since δ<1/2. If instead 1≤|kmin|≪|k|∼|kmax| is valid, then
supk≠0(∑k1+k2=k1⟨φij⟩2−2δ)≲supk≠0(1|k|2−2δ∑|kmin|≪|k|1|kmin|2−2δ)≲1, |
again due to δ<1/2. This concludes the argument for (2.11) and the whole proof.
Remark 3.2. It is important to recognize that we used s>0 only in the argument for (2.10), while the proofs for (2.8), (2.9), (2.11), and (2.12) required only that s≥0.
For the Kawahara Cauchy problem (1.2), we proceed in similar fashion to the previous section and we reformulate it in a form amenable to the framework we want to implement. First, we notice that its smooth solutions satisfy
∫T∂tu(t,x)dx=0, |
which implies
∫Tu(t,x)dx=∫Tu(0,x)dx=∫Tu0(x)dx:=μ. | (4.1) |
If we take
(v,v0):=(u−μ,u0−μ), |
then (1.2) becomes
{∂tv+β∂3xv−∂5xv+2μ∂xv=−∂x(v2),v(0,x)=v0(x), | (4.2) |
and now, due to (4.1), one has
∫Tv(t,x)dx=0. | (4.3) |
This will turn out to be important moving forward.
Next, by turning to the Fourier series coefficients (2.1) and (2.2) for v0 and v, respectively, we obtain the equivalent form of (4.2) as an infinite coupled system of ordinary differential equations:
{∂tvk+i(−βk3−k5+2μk)vk=−ik∑k1,k2∈Zk1+k2=kvk1vk2vk(0)=v0k. |
In the end, we let
wk(t):=eit(−βk3−k5+2μk)vk(t) |
to rewrite the previous system as
{∂twk=−ik∑k1,k2∈Zk1+k2=keitϕwk1wk2,wk(0)=v0k, |
where
ϕ=ϕ(k,k1,k2):=−βk3−k5+2μk+βk31+k51−2μk1+βk32+k52−2μk. |
When k=k1+k2, we have the following important factorization:
ϕ=−kk1k2(5(k21+k22+k1k2)+3β). | (4.4) |
At this moment, we observe that (4.3) implies v0(t)≡0 and, hence, w0(t)≡0. This allows us to remove from the system the equation corresponding to k=0. Moreover, when k≠0, we see that (4.4) implies
ϕ=0⟺k1=0 or k2=0. |
Thus, a working version for the Cauchy problem (1.2) is represented by
{∂twk=−ik∑k1+k2=kϕ≠0eitϕwk1wk2,∀k≠0,wk(0)=v0k, | (4.5) |
for which we argue that Theorem 2.1 holds true when s>1/4. This clearly proves Theorem 1.2.
Proposition 4.1. Theorem 2.1 is valid for the Cauchy problem (4.5) when s>1/4.
Proof. All three estimates in Theorem 2.1 (i.e., (2.4)–(2.6)) to be verified share the generic profile
‖∑k1+k2=ka(k,k1,k2)yk1zk2‖l2s3≲‖y‖l2s1‖z‖l2s2. |
Using duality, this bound can be seen to be the consequence of
supk(∑k1+k2=ka2(k,k1,k2)⟨k⟩2s3⟨k1⟩2s1⟨k2⟩2s2)≲1. | (4.6) |
It is important to note that, due to (4.4), β∈{−1,0,1}, and (4.5), we can further assume above that, in our context, none of k, k1, and k2 are equal to 0.
Here, as in the argument for Proposition 3.1, we rely on the notation (3.8), the bound (3.9), and the complementary scenarios (3.10). We start by proving (2.4), for which
a(k,k1,k2)=|k|⟨ϕ⟩1/2,s1=s2=s3=s. |
Since s≥0, the estimate (3.9) implies
⟨k⟩2s⟨k1⟩2s⟨k2⟩2s≲1 |
and, hence, (4.6) holds true if we show
supk(∑k1+k2=kϕ≠0|k|2⟨ϕ⟩)≲1. | (4.7) |
If 1≤|k|≲|kmax|∼|kmin|, then we deduce from (4.4) that |ϕ|∼|k||kmax|4 and, hence,
∑k1+k2=kϕ≠0|k|2⟨ϕ⟩≲∑|k|≤|kmax|1|kmax|3≲1, |
which yields (4.7).
If 1≤|kmin|≪|k|∼|kmax|, then (4.4) implies |ϕ|∼|kmin||k|4 and, thus,
∑k1+k2=kϕ≠0|k|2⟨ϕ⟩≲∑1≤|kmin|≪|k|1|kmin||k|2≲ln|k||k|2≲1, |
which also yields (4.7).
Next, we turn to the argument for (2.5), which corresponds to (4.6) with
a(k,k1,k2)=|k|⟨ϕ⟩1−δ,s1=s3=s−2,s2=s. |
Like above, on the basis of s≥0 and (3.9), we derive
|k|2⟨k⟩2s−4⟨k1⟩2s−4⟨k2⟩2s≲⟨kmax⟩4⟨k⟩2, |
and, hence, reduce the proof of (4.6) to the one for
supk(∑k1+k2=kϕ≠0⟨kmax⟩4⟨k⟩2⟨ϕ⟩2−2δ)≲1. | (4.8) |
If 1≤|k|≲|kmax|∼|kmin|, then |ϕ|∼|k||kmax|4 and, by taking δ<3/8, we deduce
∑k1+k2=kϕ≠0⟨kmax⟩4⟨k⟩2⟨ϕ⟩2−2δ≲∑|k|≤|kmax|1|kmax|4−8δ|k|4−2δ≲1|k|4−2δ≲1, |
which yields (4.8).
If 1≤|kmin|≪|k|∼|kmax|, then |ϕ|∼|kmin||k|4 and, now using only that δ<1/2, we infer
∑k1+k2=kϕ≠0⟨kmax⟩4⟨k⟩2⟨ϕ⟩2−2δ≲∑1≤|kmin|≪|k|1|kmin|2−2δ|k|6−8δ≲1|k|6−8δ≲1, |
which also yields (4.8).
Finally, we get to prove (2.6), for which
a(k,k1,k2)=|k|,s3=s−2,s1=s2=s. |
Thus, due to the symmetry in this case with respect to the indices 1 and 2, (4.6) is valid if we show
supk(∑k1+k2=kϕ≠0|k|2⟨k⟩2s−4⟨kmax⟩2s⟨kmin⟩2s)≲1. | (4.9) |
If 1≤|k|≲|kmax|∼|kmin|, then, as s>1/4, we deduce
∑k1+k2=kϕ≠0|k|2⟨k⟩2s−4⟨kmax⟩2s⟨kmin⟩2s≲∑|k|≤|kmax|1|kmax|2s+1/2|k|3/2≲1|k|3/2≲1, |
which implies (4.9).
If 1≤|kmin|≪|k|∼|kmax|, then, since s≥0, we infer
∑k1+k2=kϕ≠0|k|2⟨k⟩2s−4⟨kmax⟩2s⟨kmin⟩2s≲∑1≤|kmin|≪|k|1|kmin|2s+2≲1, |
which also implies (4.9).
Remark 4.2. It is important to note that in the previous proposition we used s>1/4 only in the proof of (2.6), when 1≤|k|≲|kmax|∼|kmin|. All the other arguments simply required that s≥0.
Remark 4.3. It seems that the restriction s>1/4 is essential in the sense that, by switching from s−2 to s−σ for some other σ>0, one encounters it again in the same scenario. As suggested by one of the referees, this is likely one of the shortcomings of this methodology, as (2.6) is particularly not responsive to the presence of dispersion in the original equation.
The author declares he has not used Artificial Intelligence (AI) tools in the creation of this article.
The author is deeply thankful to the careful referees for comments and suggestions which significantly improved the quality of this paper. In fact, it was a comment made by one of the reviewers that led the author to improve the UWP for the Kawahara problem from the range s>1/2 to the range s>1/4.
The author declares there is no conflict of interest.
[1] |
S. K. Turitsyn, Blow-up in the Boussinesq equation, Phys. Rev. E: Stat. Nonlinear Biol. Soft Matter Phys., 47 (1993), R796–R799. https://doi.org/10.1103/PhysRevE.47.R796 doi: 10.1103/PhysRevE.47.R796
![]() |
[2] |
F. Falk, E. W. Laedke, K. H. Spatschek, Stability of solitary-wave pulses in shape-memory alloys, Phys. Rev. B: Condens. Matter Mater. Phys., 36 (1987), 3031–3041. https://doi.org/10.1103/PhysRevB.36.3031 doi: 10.1103/PhysRevB.36.3031
![]() |
[3] | J. V. Boussinesq, Theory of waves and vortices propagating along a horizontal rectangular channel, communicating to the liquid in the channel generally similar velocities of the bottom surface, J. Math. Pures Appl., 17 (1872), 55–108. |
[4] |
V. G. Makhankov, Dynamics of classical solitons (in nonintegrable systems), Phys. Rep., 35 (1978), 1–128. https://doi.org/10.1016/0370-1573(78)90074-1 doi: 10.1016/0370-1573(78)90074-1
![]() |
[5] |
T. Kakutani, H. Ono, Weak non-linear hydromagnetic waves in a cold collision-free plasma, J. Phys. Soc. Jpn., 26 (1969), 1305–1318. https://doi.org/10.1143/JPSJ.26.1305 doi: 10.1143/JPSJ.26.1305
![]() |
[6] | H. Hasimoto, Water waves-their dispersion and steepening, Kagaku, 40 (1970), 401–408. |
[7] | D. Lannes, The Water Waves Problem: Mathematical Analysis and Asymptotics, Mathematical Surveys and Monographs, American Mathematical Society, Providence, RI, 2013. https://doi.org/10.1090/surv/188 |
[8] |
Y. F. Fang, M. G. Grillakis, Existence and uniqueness for Boussinesq type equations on a circle, Commun. Partial Differ. Equations, 21 (1996), 1253–1277. https://doi.org/10.1080/03605309608821225 doi: 10.1080/03605309608821225
![]() |
[9] |
L. G. Farah, M. Scialom, On the periodic "good" Boussinesq equation, Proc. Am. Math. Soc., 138 (2010), 953–964. https://doi.org/10.1090/S0002-9939-09-10142-9 doi: 10.1090/S0002-9939-09-10142-9
![]() |
[10] |
S. Oh, A. Stefanov, Improved local well-posedness for the periodic "good" Boussinesq equation, J. Differ. Equations, 254 (2013), 4047–4065. https://doi.org/10.1016/j.jde.2013.02.006 doi: 10.1016/j.jde.2013.02.006
![]() |
[11] |
N. Kishimoto, Sharp local well-posedness for the "good" Boussinesq equation, J. Differ. Equations, 254 (2013), 2393–2433. https://doi.org/10.1016/j.jde.2012.12.008 doi: 10.1016/j.jde.2012.12.008
![]() |
[12] |
D. A. Geba, A. A. Himonas, D. Karapetyan, Ill-posedness results for generalized Boussinesq equations, Nonlinear Anal. Theory Methods Appl., 95 (2014), 404–413. https://doi.org/10.1016/j.na.2013.09.017 doi: 10.1016/j.na.2013.09.017
![]() |
[13] |
M. Okamoto, Norm inflation for the generalized Boussinesq and Kawahara equations, Nonlinear Anal., 157 (2017), 44–61. https://doi.org/10.1016/j.na.2017.03.011 doi: 10.1016/j.na.2017.03.011
![]() |
[14] |
H. Hirayama, Local well-posedness for the periodic higher order KdV type equations, Nonlinear Differ. Equations Appl., 19 (2012), 677–693. https://doi.org/10.1007/s00030-011-0147-9 doi: 10.1007/s00030-011-0147-9
![]() |
[15] |
T. Kato, Low regularity well-posedness for the periodic Kawahara equation, Differ. Integr. Equations, 25 (2012), 1011–1036. https://doi.org/10.57262/die/1356012249 doi: 10.57262/die/1356012249
![]() |
[16] |
T. Kato, On nonlinear Schrödinger equations, Ⅱ. Hs-solutions and unconditional well-posedness, J. Anal. Math., 67 (1995), 281–306. https://doi.org/10.1007/BF02787794 doi: 10.1007/BF02787794
![]() |
[17] |
S. Kwon, T. Oh, H. Yoon, Normal form approach to unconditional well-posedness of nonlinear dispersive PDEs on the real line, Ann. Fac. Sci. Toulouse Math., 29 (2020), 649–720. https://doi.org/10.5802/afst.1643 doi: 10.5802/afst.1643
![]() |
[18] | N. Kishimoto, Unconditional uniqueness of solutions for nonlinear dispersive equations, preprint, arXiv: 1911.04349. |
[19] |
A. V. Babin, A. A. Ilyin, E. S. Titi, On the regularization mechanism for the periodic Korteweg-de Vries equation, Comm. Pure Appl. Math., 64 (2011), 591–648. https://doi.org/10.1002/cpa.20356 doi: 10.1002/cpa.20356
![]() |
[20] |
S. Kwon, T. Oh, On unconditional well-posedness of modified KdV, Int. Math. Res. Not., 2012 (2012), 3509–3534. https://doi.org/10.1093/imrn/rnr156 doi: 10.1093/imrn/rnr156
![]() |
[21] |
Z. Guo, S. Kwon, T. Oh, Poincaré-Dulac normal form reduction for unconditional well-posedness of the periodic cubic NLS, Commun. Math. Phys., 322 (2013), 19–48. https://doi.org/10.1007/s00220-013-1755-5 doi: 10.1007/s00220-013-1755-5
![]() |
[22] |
N. Kishimoto, Unconditional uniqueness for the periodic Benjamin-Ono equation by normal form approach, J. Math. Anal. Appl., 514 (2022), 126309. https://doi.org/10.1016/j.jmaa.2022.126309 doi: 10.1016/j.jmaa.2022.126309
![]() |
[23] | T. Kato, K. Tsugawa, Cancellation properties and unconditional well-posedness for the fifth order KdV type equations with periodic boundary condition, preprint, arXiv: 2308.07190. |
[24] |
L. G. Farah, Local solutions in Sobolev spaces and unconditional well-posedness for the generalized Boussinesq equation, Commun. Pure Appl. Anal., 8 (2009), 1521–1539. https://doi.org/10.3934/cpaa.2009.8.1521 doi: 10.3934/cpaa.2009.8.1521
![]() |
[25] |
D. A. Geba, B. Lin, Unconditional well-posedness for the Kawahara equation, J. Math. Anal. Appl., 502 (2021), 125282. https://doi.org/10.1016/j.jmaa.2021.125282 doi: 10.1016/j.jmaa.2021.125282
![]() |
[26] |
N. Kishimoto, Unconditional local well-posedness for periodic NLS, J. Differ. Equations, 274 (2021), 766–787. https://doi.org/10.1016/j.jde.2020.10.025 doi: 10.1016/j.jde.2020.10.025
![]() |
[27] |
N. Kishimoto, Unconditional uniqueness for the periodic modified Benjamin-Ono equation by normal form approach, Int. Math. Res. Not., 2022 (2022), 12180–12219. https://doi.org/10.1093/imrn/rnab079 doi: 10.1093/imrn/rnab079
![]() |
[28] |
N. Kishimoto, K. Tsugawa, Local well-posedness for quadratic nonlinear Schrödinger equations and the "good" Boussinesq equation, Differ. Integr. Equations, 23 (2010), 463–493. https://doi.org/10.57262/die/1356019307 doi: 10.57262/die/1356019307
![]() |