This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Pattern transformation in higher-order lumps of the Kadomtsev-Petviashvili I equation

Bo Yang and Jianke Yang Department of Mathematics and Statistics, University of Vermont, Burlington, VT 05405, U.S.A.
Abstract

Pattern formation in higher-order lumps of the Kadomtsev-Petviashvili I equation at large time is analytically studied. For a broad class of these higher-order lumps, we show that two types of solution patterns appear at large time. The first type of patterns comprise fundamental lumps arranged in triangular shapes, which are described analytically by root structures of the Yablonskii–Vorob’ev polynomials. As time evolves from large negative to large positive, this triangular pattern reverses itself along the xx-direction. The second type of patterns comprise fundamental lumps arranged in non-triangular shapes in the outer region, which are described analytically by nonzero-root structures of the Wronskian–Hermit polynomials, together with possible fundamental lumps arranged in triangular shapes in the inner region, which are described analytically by root structures of the Yablonskii–Vorob’ev polynomials. When time evolves from large negative to large positive, the non-triangular pattern in the outer region switches its xx and yy directions, while the triangular pattern in the inner region, if it arises, reverses its direction along the xx-axis. Our predicted patterns at large time are compared to true solutions, and excellent agreement is observed.

I Introduction

The Kadomtsev-Petviashvili (KP) equation was derived as a two-dimensional generalization of the Korteweg-de Vries equation for the evolution of weakly nonlinear plasma waves and shallow water waves KP1970 ; Ablowitz1979 . In the water wave context, this equation reads Ablowitz1979

[2ft+3ffx+(13T)fxxx]x+fyy=0,\left[2f_{t}+3ff_{x}+(\frac{1}{3}-T)f_{xxx}\right]_{x}+f_{yy}=0, (1)

where the spatial coordinate xx is relative to a certain moving frame, f(x,y,t)f(x,y,t) represents the water surface elevation, and TT is a dimensionless surface tension parameter. If the surface tension is large, i.e., T>1/3T>1/3, which corresponds to very thin sheets of water, this equation is called KP-I. In this case, rescaling variables by

y=y^3(T13),t=2t^T13,f=2(T13)uy=\frac{\hat{y}}{\sqrt{3(T-\frac{1}{3}})},\quad t=-\frac{2\hat{t}}{T-\frac{1}{3}},\quad f=-2(T-\frac{1}{3})u (2)

and dropping the hats, this equation becomes

(ut+6uux+uxxx)x3uyy=0.\left(u_{t}+6uu_{x}+u_{xxx}\right)_{x}-3u_{yy}=0. (3)

Note that the KP-I equation also arises in other branches of physics, such as nonlinear optics Pelinovsky1995 and Bose-Einstein condensates Tsuchiya_BEC2008 .

The KP-I equation (3) is solvable by the inverse scattering transform Zakharov_book ; Ablowitz_book . It admits stable fundamental lump solutions that are bounded rational functions decaying in all spatial directions Petviashvili1976 ; Manakov1977 ; Ablowitz_Satsuma1979 . These lumps are the counterparts of solitons in the Korteweg-de Vries equation. In the water wave context, these lumps physically correspond to dips on the water surface due to the negative sign in the ff scaling above. The KP-I equation also admits a broad class of rational solutions that describe the interactions of these lumps. If individual lumps have distinct asymptotic velocities, then they would pass through each other without change in velocities or phases Manakov1977 ; Ablowitz_Satsuma1979 . But if they have the same asymptotic velocities, they would undergo novel anomalous scattering, where the lumps would separate from each other in new spatial directions that are very different from their original incoming directions Peli93b ; Ablowitz97 ; Ablowitz2000 . In this article, we are concerned with this latter type of solutions, which we will call higher-order lumps (they are also called multi-pole lumps in the literature Ablowitz97 ; Ablowitz2000 ).

Analytical expressions of higher-order lumps have been derived by a wide variety of methods before Peli93a ; Peli93b ; Ablowitz97 ; Peli98 ; Ablowitz2000 ; Dubard2010 ; Dubard2013 ; Ma2015 ; Chen2016 ; ClarksonDowie2017 ; Gaillard2018 ; Chang2018 . Gorshkov, et al. Peli93b reported a second-order lump solution that describes the interaction and anomalous scattering of two lumps. Ablowitz et al. Ablowitz97 ; Ablowitz2000 derived higher-order lumps by the inverse scattering transform and Darboux transformation, and reproduced the solution in Peli93b as a special case. They also showed that when |t||t|\to\infty, these higher-order lumps generically split into a certain number of fundamental lumps, whose relative spatial separations grow in proportion to |t|q|t|^{q}, where 13q12\frac{1}{3}\leq q\leq\frac{1}{2}. In addition, some new lump patterns such as squares at large time were reported. Pelinovsky and Stepanyants Peli93a reported a class of higher-order lump solutions that are stationary in a moving frame. Pelinovsky Peli98 studied rational solutions of the KP hierarchy and linked them to the dynamics of the Calogero–Moser hierarchy (but his Wronskian-form solutions for KP-I were not made real-valued and thus were not physical solutions). Dubard et al. Dubard2010 ; Dubard2013 constructed a class of higher-order KP-I lump solutions from higher-order rogue waves of the nonlinear Schrödinger equation, and graphically showed that such second- and third-order lump solutions split into triangles of fundamental lumps when |t||t|\to\infty. Chen et al. Chen2016 considered a certain class of higher-order lump solutions, and graphically observed that these solutions evolve from a vertical line of fundamental lumps to a horizontal line of fundamental lumps in the (x,y)(x,y) plane when time goes from negative infinity to infinity. They also predicted the locations of fundamental lumps inside the solution complex at t=0t=0 by roots of certain polynomial equations; but such polynomial equations were not justified. Clarkson and Dowie ClarksonDowie2017 derived a second-order lump solution which incorporates the ones in Dubard2010 ; Dubard2013 ; Peli93b as special cases. Gaillard Gaillard2018 studied a special class of higher-order lump solutions and reported lump patterns such as triangles and pentagons at t=0t=0 when some internal parameters in such solutions get large. Chang Chang2018 studied the large-time asymptotics of higher-order lumps and showed that, for some special solutions, all lumps are located on a vertical line in the (x,y)(x,y) plane at large negative time but rotate to a horizontal line at large positive time. Ma Ma2015 derived a fundamental lump solution which contains more free parameters; but that solution can be made equivalent to the original lump solution as reported in Manakov1977 ; Ablowitz_Satsuma1979 . We note by passing that non-rational KP-I solutions in the form of a linear periodic chain of lumps, and those that describe the resonant collision between lumps and line solitons, have also been reported recently Zakharov2021 ; He_KPI_rogue_lump .

In this article, we study pattern formation in higher-order lumps of the KP-I equation (3). This work is motivated by our earlier work on pattern formation of rogue waves in various integrable systems YangYang21a ; YangYang21b , where we showed that universal rogue patterns appear when one of the internal parameters in rogue waves gets large, and those rogue patterns are analytically described by root structures of the Yablonskii–Vorob’ev polynomial hierarchy. For higher-order lumps of the KP-I equation, however, we will focus on their pattern formation at large time rather than at large parameters. In particular, we are interested to know how their patterns at large positive time relate to their patterns at large negative time. For a broad class of higher-order lump solutions, we will show that two types of lump patterns appear at large time. The first type of patterns comprise fundamental lumps arranged in triangular shapes, which are described analytically by root structures of the Yablonskii–Vorob’ev polynomials. As time evolves from large negative to large positive, this triangular pattern reverses itself along the xx-direction. The second type of patterns comprise fundamental lumps arranged in non-triangular shapes in the outer region, which are described analytically by nonzero-root structures of the Wronskian–Hermit polynomials, together with possible fundamental lumps arranged in triangular shapes in the inner region, which are described analytically by root structures of the Yablonskii–Vorob’ev polynomials. When time evolves from large negative to large positive, the non-triangular pattern in the outer region switches its xx and yy directions, plus some rescaling along each direction, while the triangular pattern in the inner region, if it arises, reverses its direction along the xx-axis. These dramatic pattern transformations with the elapse of time are fascinating. We have also compared these predicted patterns with true solutions, and excellent agreement is observed.

This paper is organized as follows. In Sec. 2, we present general higher-order lump solutions in the KP-I equation through Schur polynomials, and introduce Yablonskii–Vorob’ev and Wronskian–Hermit polynomials. In Sec. 3, we present our main analytical results on solution patterns at large time, and explain how these patterns transform from large negative time to large positive time. In Sec. 4, we illustrate our pattern predictions and compare them with true solutions. In Sec. 5, we provide proofs for our analytical results in Sec. 3. The last section summarizes our results, together with some discussions. In the Appendix, a brief derivation of our general higher-order lump solutions in Sec. 2 is given.

II Preliminaries

The KP equation (3) admits three important invariances. The first one is that it is invariant when (x,t)(x,t)(x,t)\to(-x,-t). This invariance is important because it shows that KP-solution patterns are reversible in time (albeit with a sign switch in xx). In earlier works Chen2016 ; Chang2018 , the authors showed that certain higher-order KP lumps evolve from a vertical line of fundamental lumps to a horizontal line of fundamental lumps in the (x,y)(x,y) plane when time goes from negative infinity to infinity. The above invariance indicates that a reverse pattern transformation could also occur, i.e., those higher-order KP lumps can also evolve from a horizontal line to a vertical line when time goes from negative infinity to infinity.

The second invariance of the KP equation (3) is the Galilean invariance Weiss1985 ; Chen2016 , i.e., when

(x,y,t)(x+2ky+12k2t,y+12kt,t),(x,y,t)\to(x+2ky+12k^{2}t,\,y+12kt,\,t), (4)

the KP solution u(x,y,t)u(x,y,t) remains a solution. Here, kk is an arbitrary real constant. This invariance indicates that, if the overall solution complex has a yy-direction velocity 12k12k, then we can apply this invariance to remove that yy-direction velocity. In doing so, the solution pattern in the (x,y)(x,y) plane would change as well through a linear transformation of shear type. This Galilean invariance is important, because it allows us to remove the overall yy-direction velocity in a higher-order lump solution. More will be said on it later in this section.

The third invariance of the KP equation is scaling invariance, i.e., when

(x,y,t,u)(αx,α2y,α3t,α2u),(x,y,t,u)\to(\alpha x,\alpha^{2}y,\alpha^{3}t,\alpha^{-2}u), (5)

the KP equation remains invariant. Here, α\alpha is any nonzero real constant. This invariance is useful since, when combined with the Galilean invariance above, it allows us to normalize the spectral parameter in the KP-lump solutions to be unity without any loss of generality. This we will do in Sec. 3.

II.1 Explicit expressions of higher-order lumps

In this paper, we consider pattern formation of higher-order lumps in the KP-I equation (3). General higher-order lump solutions have been derived by Ablowitz et al. Ablowitz2000 through Darboux transformation. Their solutions were given through determinants whose matrix elements involve differential operators with respect to the spectral parameter. For our analysis, those solution expressions are not explicit enough. Thus, we have derived these higher-order lumps again by the bilinear method. To present our solutions, we first introduce elementary Schur polynomials Sk(𝒙)S_{k}(\mbox{\boldmath$x$}) with x=(x1,x2,)\emph{{x}}=\left(x_{1},x_{2},\ldots\right), which are defined by the generating function

n=0Sn(𝒙)ϵn=exp(n=1xnϵn).\sum_{n=0}^{\infty}S_{n}(\mbox{\boldmath$x$})\epsilon^{n}=\exp\left(\sum_{n=1}^{\infty}x_{n}\epsilon^{n}\right). (6)

More explicitly,

S0(x)=1,S1(𝒙)=x1,S2(𝒙)=12x12+x2,,Sn(𝒙)=l1+2l2++mlm=n(j=1mxjljlj!).S_{0}(\emph{{x}})=1,\quad S_{1}(\mbox{\boldmath$x$})=x_{1},\quad S_{2}(\mbox{\boldmath$x$})=\frac{1}{2}x_{1}^{2}+x_{2},\quad\cdots,\quad S_{n}(\mbox{\boldmath$x$})=\sum_{l_{1}+2l_{2}+\cdots+ml_{m}=n}\left(\ \prod_{j=1}^{m}\frac{x_{j}^{l_{j}}}{l_{j}!}\right).

Under these notations, our general higher-order KP-I lumps are given by the following theorem.

Theorem 1General higher-order lumps of the KP-I equation (3) are

uΛ(x,y,t)=2x2lnσ,\displaystyle u_{\Lambda}(x,y,t)=2\partial_{x}^{2}\ln\sigma, (7)

where

σ(x,y,t)=det1i,jN(mij),\sigma(x,y,t)=\det_{1\leq i,j\leq N}\left(m_{ij}\right), (8)
mi,j=ν=0min(ni,nj)[|p|2(p+p)2]νSniν(x++νs+ai)Snjν[(x+)+νs+aj],m_{i,j}=\sum_{\nu=0}^{\min(n_{i},n_{j})}\left[\frac{|p|^{2}}{(p+p^{*})^{2}}\right]^{\nu}\hskip 1.70709ptS_{n_{i}-\nu}(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}}_{i})\hskip 1.70709ptS_{n_{j}-\nu}[(\textbf{\emph{x}}^{+})^{*}+\nu\textbf{\emph{s}}^{*}+\textbf{\emph{a}}_{j}^{*}], (9)

NN is an arbitrary positive integer, Λ(n1,n2,nN)\Lambda\equiv(n_{1},n_{2},\cdots n_{N}) is a vector of arbitrary positive integers, pp is an arbitrary non-imaginary complex number, the asterisk ‘*’ represents complex conjugation, the vector x+=(x1+,x2+,)\textbf{\emph{x}}^{+}=\left(x_{1}^{+},x_{2}^{+},\cdots\right) is defined by

xk+=p1k!x+p22kk!iy+p33kk!(4)t,\displaystyle x_{k}^{+}=p\frac{1}{k!}x+p^{2}\frac{2^{k}}{k!}\textrm{i}y+p^{3}\frac{3^{k}}{k!}(-4)t, (10)

the vector s=(s1,s2,)\textbf{\emph{s}}=(s_{1},s_{2},\cdots) is defined through the expansion

ln[1κ(p+p)(eκ1peκ+p)]=j=1sjκj,\displaystyle\ln\left[\frac{1}{\kappa}\left(p+p^{*}\right)\left(\frac{e^{\kappa}-1}{p\hskip 1.42271pte^{\kappa}+p^{*}}\right)\right]=\sum_{j=1}^{\infty}s_{j}\hskip 1.42271pt\kappa^{j}, (11)

vectors ai\textbf{\emph{a}}_{i} are

ai=(ai,1,ai,2,,ai,ni),\textbf{\emph{a}}_{i}=\left(a_{i,1},a_{i,2},\cdots,a_{i,n_{i}}\right), (12)

and ai,j(1iN,1jni)a_{i,j}\hskip 1.42271pt(1\leq i\leq N,1\leq j\leq n_{i}) are free complex constants.

The proof of this theorem will be given in the Appendix.

Remark 1. In this theorem, positive integers (n1,n2,nN)(n_{1},n_{2},\cdots n_{N}) do not have to be distinct if their corresponding vectors ai\textbf{\emph{a}}_{i} are different. In such cases, by first rewriting the σ\sigma determinant (8) as a larger determinant as was done in Ref. OhtaJY2012 , then linking Schur polynomials with different ai\textbf{\emph{a}}_{i} vectors in that larger determinant by relations similar to Eq. (167) in Ref. YangYang3wave , and finally applying row operations and parameter redefinitions to the resulting determinant, we can show that this σ\sigma determinant (8) with non-distinct integers (n1,n2,nN)(n_{1},n_{2},\cdots n_{N}) can be reduced to one where the new integers (n^1,n^2,n^N)(\hat{n}_{1},\hat{n}_{2},\cdots\hat{n}_{N}) become distinct. Thus, in this paper, we will require positive integers (n1,n2,nN)(n_{1},n_{2},\cdots n_{N}) to be distinct without loss of generality. In this case, we will also arrange them in the ascending order, i.e., n1<n2<<nNn_{1}<n_{2}<\cdots<n_{N}.

Remark 2. The higher-order lumps in Theorem 1 contain free complex parameters pp and ai\textbf{\emph{a}}_{i} (1iN1\leq i\leq N), totaling 1+n1+n2+nN1+n_{1}+n_{2}+\cdots n_{N}. However, using techniques similar to that outlined in Remark 1, we can show that N(N1)/2N(N-1)/2 of those parameters in {ai}\{\textbf{\emph{a}}_{i}\} can be eliminated. Thus, the number of free complex parameters in these higher-order lumps can be reduced to 1+ρ1+\rho, where

ρ=i=1NniN(N1)2.\rho=\sum_{i=1}^{N}n_{i}-\frac{N(N-1)}{2}. (13)

This number of free parameters matches that given in Ref. Ablowitz2000 for solutions produced by Darboux transformation. In fact, from the derivation of Theorem 1 in the Appendix, we can see that our higher-order lumps given in this theorem by the bilinear method are equivalent to those derived in Ablowitz2000 by Darboux transformation, except that our expressions are more explicit.

Remark 3. The fundamental lump can be derived by taking N=1N=1 and n1=1n_{1}=1 in Eq. (8). Through a shift of the (x,y)(x,y) axes, we can normalize a1,1=0a_{1,1}=0. Then, the resulting σ1(x,y,t)\sigma_{1}(x,y,t) function can be reduced to

σ1=|x+2ipy12p2t|2+1(p+p)2\displaystyle\sigma_{1}=\left|x+2\textrm{i}py-12p^{2}t\right|^{2}+\frac{1}{(p+p^{*})^{2}}
=[x2piy12(pr2pi2)t]2+[2pr(y12pit)]2+14pr2,\displaystyle\hskip 8.5359pt=\left[x-2p_{i}\hskip 0.85355pty-12(p_{r}^{2}-p_{i}^{2})\hskip 0.85355ptt\right]^{2}+\left[2p_{r}(y-12p_{i}\hskip 0.85355ptt)\right]^{2}+\frac{1}{4p_{r}^{2}}, (14)

where prp_{r} and pip_{i} are the real and imaginary parts of the spectral parameter pp. The corresponding solution u1(x,y,t)u_{1}(x,y,t) through Eq. (7) moves at xx-direction velocity of 12|p|212|p|^{2} and yy-direction velocity of 12pi12p_{i}. By applying the Galilean invariance (4) with k=pik=p_{i}, we can remove the yy-direction velocity 12pi12p_{i} and reduce σ1(x,y,t)\sigma_{1}(x,y,t) to

σ1=(x12pr2t)2+(2pry)2+14pr2.\displaystyle\sigma_{1}=\left(x-12p_{r}^{2}t\right)^{2}+\left(2p_{r}y\right)^{2}+\frac{1}{4p_{r}^{2}}. (15)

This means that, under Galilean invariance, we can take pp in the original fundamental lump to be purely real without loss of generality. Then, by utilizing the scaling invariance (5) with α=pr\alpha=p_{r}, we can further normalize prp_{r} in the above σ1\sigma_{1} to be unity. The final simplified fundamental-lump expression is

u1(x,y,t)=2x2ln[(x12t)2+4y2+14].u_{1}(x,y,t)=2\partial_{x}^{2}\ln\left[\left(x-12t\right)^{2}+4y^{2}+\frac{1}{4}\right]. (16)

This is a moving single lump with peak amplitude 1616, which is attained at the spatial location of (x,y)=(12t,0)(x,y)=(12t,\hskip 1.42271pt0).

Remark 4. In the general higher-order lump of Theorem 1, the whole solution complex moves at xx-direction velocity 12|p|212|p|^{2} and yy-direction velocity 12pi12p_{i}, plus some possible slower motion relative to that moving frame. In this general case, we can also use the Galilean invariance (4) to remove the yy-direction velocity 12pi12p_{i} of the complex, i.e., pip_{i} can be made to be zero. In addition, we can use the scaling invariance (5) to normalize prp_{r} to unity. Thus, without any loss of generality, we can choose pp in the higher-order lump solution of Theorem 1 to be equal to one. For this reason, we will set p=1p=1 in the remainder of this paper.

Remark 5. In Peli93a , a class of higher-order lump solutions that are stationary in a moving frame was reported. Those special solutions satisfy the Boussinesq equation. Thus, they are special cases of Boussinesq rogue waves YangYangBoussi . Those stationary higher-order lumps are also special cases of our solutions in Theorem 1 when the index vector (n1,n2,nN)(n_{1},n_{2},\cdots n_{N}) and internal parameters {ai}\{\textbf{\emph{a}}_{i}\} are properly chosen. Indeed, rational solutions in Theorem 1 would be stationary if the σ\sigma function in (8) satisfies the dimension-reduction condition σtVσx=0\sigma_{t}-V\sigma_{x}=0, where VV is the velocity of the moving frame along the xx-direction. In the bilinear derivation of Boussinesq rogue waves YangYangBoussi , one needs to solve the bilinear τ\tau equation of KP-I, together with this τ\tau’s dimension reduction condition τx33τx1=Cτ\tau_{x_{3}}-3\tau_{x_{1}}=C\tau, where x1x_{1} is proportional to xx, x3x_{3} proportional to tt, and CC is a constant. Since this τ\tau function turns out to be equal to σ\sigma multiplying an exponential of a linear function of xx and tt YangYangBoussi , we see that τ\tau’s dimension reduction condition is equivalent to σ\sigma’s dimension reduction condition after proper variable scalings. This means that constraints from τ\tau’s dimension-reduction condition can be borrowed over and imposed on solutions of Theorem 1 in order to obtain stationary higher-order KP lumps. One of such constraints is on the index vector (n1,n2,nN)(n_{1},n_{2},\cdots n_{N}), where ni=2i1n_{i}=2i-1 must be chosen YangYangBoussi . In addition, internal parameters {ai}\{\textbf{\emph{a}}_{i}\} also need to be constrained. For a different choice of differential operators than those in Eq. (105) of the Appendix, this parameter constraint was derived in YangYangBoussi . For the present choice of differential operators in Eq. (105), this parameter constraint would be more complex. In this case, such a parameter constraint was worked out in ChenJunChao2018 for another integrable system under a different parameterization.

II.2 Yablonskii–Vorob’ev polynomials and Wronskian-Hermit polynomials

We will show in later text that patterns of certain higher-order lump solutions at large time are described by root structures of the Yablonskii–Vorob’ev polynomials and Wronskian-Hermit polynomials. Thus, these polynomials and their root structures will be introduced first.

II.2.1 Yablonskii–Vorob’ev polynomials and their root structures

Yablonskii-Vorob’ev polynomials arose in rational solutions of the second Painlevé equation (PII\mbox{P}_{\mbox{\scriptsize II}}) Yablonskii1959 ; Vorobev1965 . Later, a determinant expression for these polynomials was found in Kajiwara-Ohta1996 . Let pk(z)p_{k}(z) be polynomials defined by

k=0pk(z)ϵk=exp(zϵ43ϵ3).\sum_{k=0}^{\infty}p_{k}(z)\epsilon^{k}=\exp\left(z\epsilon-\frac{4}{3}\epsilon^{3}\right). (17)

Then, Yablonskii-Vorob’ev polynomials QN(z)Q_{N}(z) are given by the N×NN\times N determinant Kajiwara-Ohta1996

QN(z)=cN|p1(z)p0(z)p2N(z)p3(z)p2(z)p4N(z)p2N1(z)p2N2(z)pN(z)|,\displaystyle Q_{N}(z)=c_{N}\left|\begin{array}[]{cccc}p_{1}(z)&p_{0}(z)&\cdots&p_{2-N}(z)\\ p_{3}(z)&p_{2}(z)&\cdots&p_{4-N}(z)\\ \vdots&\vdots&\vdots&\vdots\\ p_{2N-1}(z)&p_{2N-2}(z)&\cdots&p_{N}(z)\end{array}\right|, (22)

where cN=j=1N(2j1)!!c_{N}=\prod_{j=1}^{N}(2j-1)!!, and pk(z)0p_{k}(z)\equiv 0 if k<0k<0. This determinant is a Wronskian since one can see from Eq. (17) that pk+1(z)=pk(z)p^{\prime}_{k+1}(z)=p_{k}(z), where the prime represents differentiation. Yablonskii-Vorob’ev polynomials are monic polynomials with integer coefficients Clarkson2003-II , and the first four of them are

Q1(z)=z,\displaystyle Q_{1}(z)=z,
Q2(z)=z3+4,\displaystyle Q_{2}(z)=z^{3}+4,
Q3(z)=z6+20z380,\displaystyle Q_{3}(z)=z^{6}+20z^{3}-80,
Q4(z)=z(z9+60z6+11200).\displaystyle Q_{4}(z)=z(z^{9}+60z^{6}+11200).

Root structures of these polynomials have been studied in Fukutani ; Taneda ; Clarkson2003-II ; Miller2014 ; Bertola2016 , and the following facts are known.

  1. 1.

    The degree of the QN(z)Q_{N}(z) polynomial is N(N+1)/2N(N+1)/2, which can be easily seen from Eq. (22).

  2. 2.

    All roots of QN(z)Q_{N}(z) are simple Fukutani . Thus, QN(z)Q_{N}(z) has N(N+1)/2N(N+1)/2 simple roots.

  3. 3.

    Zero is a root of QN(z)Q_{N}(z) if and only if N1mod3N\equiv 1\hskip 2.84544pt\mbox{mod}\hskip 2.84544pt3 Taneda .

  4. 4.

    QN(z)Q_{N}(z) can be factorized as QN(z)=zmf(ζ)Q_{N}(z)=z^{m}f(\zeta), where m=1m=1 when N1mod3N\equiv 1\hskip 2.84544pt\mbox{mod}\hskip 2.84544pt3 and m=0m=0 otherwise, ζz3\zeta\equiv z^{3}, and f(ζ)f(\zeta) is a polynomial of ζ\zeta with integer coefficients and a nonzero constant term Clarkson2003-II . This factorization shows that the root structure of QN(z)Q_{N}(z) is invariant under 120120^{\circ}-angle rotation in the complex zz plane.

  5. 5.

    Roots of QN(z)Q_{N}(z) exhibit a triangular pattern in the complex plane for all N2N\geq 2 Clarkson2003-II ; Miller2014 ; Bertola2016 . This fact is not surprising given the 120120^{\circ} rotational symmetry of QN(z)Q_{N}(z)’s root structure mentioned above.

  6. 6.

    Roots of QN(z)Q_{N}(z) are also symmetric with respect to the real-zz axis, since the coefficients of QN(z)Q_{N}(z) are real and thus complex roots appear in conjugate pairs. This conjugate symmetry, together with the 120120^{\circ} rotational symmetry, implies that one vertex of the triangular root structure of QN(z)Q_{N}(z) is on the real-zz axis.

Due to importance of these root structures to our work, we reproduce some of them in Fig. 1 for 2N52\leq N\leq 5.

Refer to caption
Figure 1: Root structures of Yablonskii-Vorob’ev polynomials QN(z)Q_{N}(z) in the complex zz plane for 2N52\leq N\leq 5.

II.2.2 Wronskian-Hermit polynomials and their root structures

Next, we introduce Wronskian-Hermit polynomials. Let qk(z)q_{k}(z) be polynomials defined by

k=0qk(z)ϵk=exp(zϵ+ϵ2).\sum_{k=0}^{\infty}q_{k}(z)\epsilon^{k}=\exp\left(z\epsilon+\epsilon^{2}\right). (23)

These qk(z)q_{k}(z) polynomials are related to Hermit polynomials through simple variable scalings. Then, for any positive integer NN and index vector Λ=(n1,n2,,nN)\Lambda=(n_{1},n_{2},\dots,n_{N}), where {ni}\{n_{i}\} are positive and distinct integers in ascending order, i.e., n1<n2<<nNn_{1}<n_{2}<\cdots<n_{N}, the Wronskian-Hermite polynomial WΛ(z)W_{\Lambda}(z) is defined as the Wronskian of qk(z)q_{k}(z) polynomials

WΛ(z)=Wron[qn1(z),qn2(z),,qnN(z)],W_{\Lambda}(z)=\mbox{Wron}\left[q_{n_{1}}(z),q_{n_{2}}(z),\ldots,q_{n_{N}}(z)\right], (24)

or equivalently,

WΛ(z)=|qn1(z)qn11(z)qn1N+1(z)qn2(z)qn21(z)qn2N+1(z)qnN(z)qnN1(z)qnNN+1(z)|,W_{\Lambda}(z)=\left|\begin{array}[]{cccc}q_{n_{1}}(z)&q_{n_{1}-1}(z)&\cdots&q_{n_{1}-N+1}(z)\\ q_{n_{2}}(z)&q_{n_{2}-1}(z)&\cdots&q_{n_{2}-N+1}(z)\\ \vdots&\vdots&\vdots&\vdots\\ q_{n_{N}}(z)&q_{n_{N}-1}(z)&\cdots&q_{n_{N}-N+1}(z)\end{array}\right|, (25)

since we can see qk+1(z)=qk(z)q_{k+1}^{\prime}(z)=q_{k}(z) from the definition (23). In the above determinant, qk(z)0q_{k}(z)\equiv 0 when k<0k<0.

Regarding root structures of Wronskian-Hermite polynomials WΛ(z)W_{\Lambda}(z), we have the following facts.

  1. 1.

    The degree of the polynomial WΛ(z)W_{\Lambda}(z) is equal to ρ\rho, where ρ\rho is given in Eq. (13). This fact can be seen from the definition (25).

  2. 2.

    The multiplicity of the zero root in WΛ(z)W_{\Lambda}(z) is equal to d(d+1)/2d(d+1)/2, where

    d=koddkeven,d=k_{odd}-k_{even}, (26)

    and koddk_{odd}, kevenk_{even} are the numbers of odd and even elements in the index vector (n1,n2,,nN)(n_{1},n_{2},\dots,n_{N}) respectively. This fact was mentioned in Felder2012 ; Garcia_Ferrero2015 and proved in Bonneux2020 . If d(d+1)/2=0d(d+1)/2=0, i.e., d=0d=0 or 1-1, then zero is not a root of WΛ(z)W_{\Lambda}(z).

  3. 3.

    The number of nonzero roots (counting multiplicity) in WΛ(z)W_{\Lambda}(z), which we denote as NWN_{W}, is

    NW=ρd(d+1)2.N_{W}=\rho-\frac{d(d+1)}{2}. (27)
  4. 4.

    The polynomial WΛ(z)W_{\Lambda}(z) can be factored as WΛ(z)=zd(d+1)/2f(ζ)W_{\Lambda}(z)=z^{d(d+1)/2}f(\zeta), where dd is given in Eq. (26), ζz2\zeta\equiv z^{2}, and f(ζ)f(\zeta) is a polynomial of ζ\zeta with real coefficients and a nonzero constant term Bonneux2020 .

  5. 5.

    If z0z_{0} is a root of WΛ(z)W_{\Lambda}(z), so are z0,z0-z_{0},z_{0}^{*} and z0-z_{0}^{*}. This quartet root symmetry can be seen from the above factorization of WΛ(z)W_{\Lambda}(z) and the fact that the coefficients of the polynomial WΛ(z)W_{\Lambda}(z) are real. As a consequence of this quartet symmetry, the root structure of WΛ(z)W_{\Lambda}(z) is non-triangular. This is a big difference from Yablonskii-Vorob’ev polynomials, which feature triangular root structures.

In addition, we have the following lemma.

Lemma 1. The Wronskian-Hermite polynomial WΛ(z)W_{\Lambda}(z) has only zero roots, i.e., NW=0N_{W}=0, if and only if (n1,n2,,nN)=(1,3,5,,2N1)(n_{1},n_{2},\dots,n_{N})=(1,3,5,\cdots,2N-1).

Proof. Since kodd+keven=Nk_{odd}+k_{even}=N, we have from Eq. (27) that

NW=i=1Nnikodd2keven(keven1).N_{W}=\sum_{i=1}^{N}n_{i}-k_{odd}^{2}-k_{even}(k_{even}-1). (28)

Since {ni}\{n_{i}\} are distinct positive integers, their smallest possible values, after reordering, are {1,3,2kodd1,2,4,,2keven}\{1,3,\cdots 2k_{odd}-1,2,4,\cdots,2k_{even}\}. Thus,

i=1Nni[1+3+(2kodd1)]+(2+4++2keven)=kodd2+keven(keven+1).\sum_{i=1}^{N}n_{i}\geq\left[1+3+\cdots(2k_{odd}-1)\right]+\left(2+4+\cdots+2k_{even}\right)=k_{odd}^{2}+k_{even}(k_{even}+1). (29)

Then,

NW2keven.N_{W}\geq 2k_{even}. (30)

For WΛ(z)W_{\Lambda}(z) to have only zero roots, NWN_{W} must be zero; so keven=0k_{even}=0, i.e., all numbers in {ni}\{n_{i}\} must be odd. In addition, for the equality in (30) to hold, all these odd and distinct numbers must be the lowest, i.e., (n1,n2,,nN)=(1,3,5,,2N1)(n_{1},n_{2},\dots,n_{N})=(1,3,5,\cdots,2N-1). This completes the proof.

This lemma tells us that, as long as Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\cdots,2N-1), the Wronskian-Hermit polynomial WΛ(z)W_{\Lambda}(z) would always have nonzero roots. This result is important to us, as we will show in later text that the presence or absence of nonzero roots in WΛ(z)W_{\Lambda}(z) will have direct consequences on the solution patterns of higher-order lumps.

On roots of Wronskian-Hermite polynomials, beside the above facts, the following conjecture has also been proposed.

Conjecture 1 Felder2012 . All roots of every Wronskian-Hermite polynomial WΛ(z)W_{\Lambda}(z) are simple, except possibly the zero root.

This conjecture will be useful, as we will show in later text that the multiplicity of a root in the Wronskian-Hermite polynomial has direct implications on the wave structure of higher-order lumps. Based on this conjecture, WΛ(z)W_{\Lambda}(z) would have NWN_{W} nonzero simple roots, where NWN_{W} is given in Eq. (27). We have checked this conjecture on a number of examples of WΛ(z)W_{\Lambda}(z), and found it to always hold.

To illustrate root structures of Wronskian-Hermite polynomials, we choose two index vectors

Λ1=(2,3,4,5),Λ2=(3,4,5,7,9).\Lambda_{1}=(2,3,4,5),\quad\Lambda_{2}=(3,4,5,7,9). (31)

The corresponding polynomials are found to be

WΛ1(z)=z816z6+120z4+7202880,\displaystyle W_{\Lambda_{1}}(z)=\frac{z^{8}-16z^{6}+120z^{4}+720}{2880}, (32)
WΛ2(z)=z6(z1212z10+180z8+672z67056z4181440z21270080)2743372800.\displaystyle W_{\Lambda_{2}}(z)=-\frac{z^{6}\left(z^{12}-12z^{10}+180z^{8}+672z^{6}-7056z^{4}-181440z^{2}-1270080\right)}{2743372800}. (33)

Root structures of these two polynomials are plotted in Fig. 2. It is seen that for the first polynomial, its root structure is rectangular and does not contain zero. For the second polynomial, its root structure is quasi-rectangular with a zero root (of multiplicity six) in the center. All nonzero roots in these two polynomials are simple, which is consistent with the earlier conjecture.

Refer to caption
Figure 2: Root structures of Wronskian-Hermite polynomials WΛ1(z)W_{\Lambda_{1}}(z) (left) and WΛ2(z)W_{\Lambda_{2}}(z) (right) in the complex plane zz, where index vectors Λ1\Lambda_{1} and Λ2\Lambda_{2} are given in Eq. (31).

III Patterns of higher-order lumps at large times

In this section, we study patterns of higher-order lumps at large times. In this study, we will set the spectral parameter p=1p=1 without loss of generality (see Remark 4 in the previous section). In this case, the constant factor in Eq. (9) simplifies to 1/4ν1/4^{\nu}. In addition, the definition (11) of the s vector reduces to

ln[2κtanh(κ2)]=k=1skκk,\displaystyle\ln\left[\frac{2}{\kappa}\tanh\left(\frac{\kappa}{2}\right)\right]=\sum_{k=1}^{\infty}s_{k}\hskip 1.42271pt\kappa^{k}, (34)

which is identical to the s vector in the earlier work OhtaJY2012 on rogue waves of the NLS equation. In particular, all odd-indexed elements sodds_{odd} of s are zero since the function on the left side of the above equation is even.

It turns out that pattern analysis of lumps depends on whether vectors of internal parameters {ai}\{\textbf{\emph{a}}_{i}\} are the same vectors (i.e., whether vector elements ai,ja_{i,j} depend on the ii index). In this paper, we only consider the case where these {ai}\{\textbf{\emph{a}}_{i}\} vectors are the same, i.e., ai=a\textbf{\emph{a}}_{i}=\textbf{\emph{a}}. In this case, through a shift of the (x,y)(x,y) axes, we can make the first element of a to be zero. Thus, our parameter choices are

ai=a=(0,a2,a3,).\textbf{\emph{a}}_{i}=\textbf{\emph{a}}=(0,a_{2},a_{3},\cdots). (35)

Under these parameters, we have two theorems on patterns of higher-order lumps at large times, depending on whether the index vector Λ\Lambda is equal to (1,3,5,,2N1)(1,3,5,\dots,2N-1).

III.1 Large-time lump patterns when Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1)

Our first theorem is for the case where the index vector Λ\Lambda is equal to (1,3,5,,2N1)(1,3,5,\dots,2N-1).

Theorem 2. If the index vector Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1), then, when |t|1|t|\gg 1, the higher-order lump solution uΛ(x,y,t)u_{\Lambda}(x,y,t) asymptotically separates into N(N+1)/2N(N+1)/2 fundamental lumps u1(xx0,yy0,t)u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t), where u1(x,y,t)u_{1}(x,y,t) is given in Eq. (16),

x0=(z0)(12t)1/3,y0=(z0)2(12t)1/3,x_{0}=\Re(z_{0})\hskip 1.42271pt(12t)^{1/3},\quad y_{0}=\frac{\Im(z_{0})}{2}(12t)^{1/3}, (36)

z0z_{0} is each of the N(N+1)/2N(N+1)/2 simple roots of the Yablonskii–Vorob’ev polynomial QN(z)Q_{N}(z), and \Re, \Im represent the real and imaginary parts of a complex number. The peak of each fundamental lump is spatially located at (x,y)=(12t+x0,y0)(x,y)=(12t+x_{0},y_{0}). The absolute error of this fundamental-lump approximation is O(|t|1/3)O(|t|^{-1/3}) when z00z_{0}\neq 0 and O(|t|1)O(|t|^{-1}) when zero is a root and z0=0z_{0}=0. Expressed mathematically, when (x,y)(x,y) is in the neighborhood of each of these fundamental lumps, i.e., (x12tx0)2+(yy0)2=O(1)(x-12t-x_{0})^{2}+(y-y_{0})^{2}=O(1), we have the following solution asymptotics for |t|1|t|\gg 1,

uΛ(x,y,t)={u1(xx0,yy0,t)+O(|t|1/3),ifz00,u1(xx0,yy0,t)+O(|t|1),ifz0=0.u_{\Lambda}(x,y,t)=\left\{\begin{array}[]{ll}u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t)+O\left(|t|^{-1/3}\right),&\mbox{if}\hskip 1.9919ptz_{0}\neq 0,\\ u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t)+O\left(|t|^{-1}\right),&\mbox{if}\hskip 1.9919ptz_{0}=0.\end{array}\right. (37)

When (x,y)(x,y) is not in the neighborhood of any of these N(N+1)/2N(N+1)/2 fundamental lumps, uΛ(x,y,t)u_{\Lambda}(x,y,t) asymptotically approaches zero as |t||t|\to\infty.

The proof of this theorem will be provided in Sec. V.

This theorem indicates that, wave patterns at large times are formed by N(N+1)/2N(N+1)/2 fundamental lumps. Relative to the moving frame of xx-direction velocity 1212, positions (x0,y0)(x_{0},y_{0}) of these fundamental lumps are just a simple linear transformation of the root structure of the Yablonskii–Vorob’ev polynomial QN(z)Q_{N}(z), i.e.,

[x0y0]=(12t)1/3[10012][(z0)(z0)].\left[\begin{array}[]{c}x_{0}\\ y_{0}\end{array}\right]=(12t)^{1/3}\left[\begin{array}[]{cc}1&0\\ 0&\frac{1}{2}\end{array}\right]\left[\begin{array}[]{c}\Re(z_{0})\\ \Im(z_{0})\end{array}\right]. (38)

Since the transformation matrix is diagonal, this transformation is simply a stretching along both horizontal and vertical directions. Recall that the Yablonskii–Vorob’ev root structure is triangular (see Fig. 1). The resulting lump pattern is then triangular as well. When t1t\gg 1, this triangular lump pattern preserves the same orientation of the original triangle of the Yablonskii–Vorob’ev root structure. But when t1t\ll-1, the triangular lump pattern would be oriented opposite of the Yablonskii–Vorob’ev root structure. Indeed, it is easy to see from Eq. (38) that, when time changes from large negative to large positive, i.e., from t-t to +t+t, their lump positions would be related as

[x0+y0+]=[x0y0].\left[\begin{array}[]{c}x_{0}^{+}\\ y_{0}^{+}\end{array}\right]=-\left[\begin{array}[]{c}x_{0}^{-}\\ y_{0}^{-}\end{array}\right]. (39)

Thus, these triangular lump patterns have reversed directions along the xx-axis (the yy-direction reversal does not matter since the pattern is symmetric in yy). This xx-direction reversal of triangular lump patterns when time changes from large negative to large positive is a dramatic pattern transformation in the KP-I equation. This phenomenon has been graphically reported in Dubard2013 on several low-order solution examples. Here, we established this fact for the general case.

Theorem 2 also indicates that, at large time, fundamental lumps in the solution complex separate from each other in proportion to |t|1/3|t|^{1/3}. This rate of separation is very slow, relative to the overall (linear) speed 1212 of the whole complex.

One more feature of Theorem 2 is that, positions (x0,y0)(x_{0},y_{0}) in Eq. (36) for individual fundamental lumps in the solution complex are independent of the solution’s internal parameters a. This means that, when |t||t|\to\infty, solutions uΛ(x,y,t)u_{\Lambda}(x,y,t) with different internal parameters a would approach the same limit solution.

III.2 Large-time lump patterns when Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1)

When Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1), the Wronskian-Hermite polynomial WΛ(z)W_{\Lambda}(z) has a zero root of multiplicity d(d+1)/2d(d+1)/2, with dd given in Eq. (26), as well as nonzero roots that are conjectured to be all simple (see Sec. II.2.2). Note that the zero root would be absent if d=0d=0 or 1-1; but nonzero roots always exist. In this case, our results on solution patterns at large time are summarized in the following theorem.

Theorem 3. Suppose the index vector Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1), and all nonzero roots of the Wronskian-Hermite polynomial WΛ(z)W_{\Lambda}(z) are simple. Then, for |t|1|t|\gg 1, the following asymptotics for the solution uΛ(x,y,t)u_{\Lambda}(x,y,t) holds.

  1. 1.

    In the outer region — the region that is O(|t|1/2)O\left(|t|^{1/2}\right) away from the wave center (x,y)=(12t,0)(x,y)=(12t,0), or (x12t)2+y2=O(|t|1/2)\sqrt{(x-12t)^{2}+y^{2}}=O\left(|t|^{1/2}\right), the higher-order lump uΛ(x,y,t)u_{\Lambda}(x,y,t) asymptotically separates into NWN_{W} fundamental lumps u1(xx0,yy0,t)u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t), where NWN_{W} is given in Eq. (27), u1(x,y,t)u_{1}(x,y,t) is given in Eq. (16),

    x0=[z0(12t)1/2]+O(1),y0=[z0(12t)1/2]2+O(1),x_{0}=\Re\left[z_{0}(-12t)^{1/2}\right]+O(1),\quad y_{0}=\frac{\Im\left[z_{0}(-12t)^{1/2}\right]}{2}+O(1), (40)

    and z0z_{0} is each of the NWN_{W} nonzero simple roots of WΛ(z)W_{\Lambda}(z). The absolute error of this fundamental-lump approximation is O(|t|1/2)O(|t|^{-1/2}). Expressed mathematically, when (x,y)(x,y) is in the neighborhood of each of these fundamental lumps, i.e., (x12tx0)2+(yy0)2=O(1)(x-12t-x_{0})^{2}+(y-y_{0})^{2}=O(1), we have the following solution asymptotics

    uΛ(x,y,t)=u1(xx0,yy0,t)+O(|t|1/2),|t|1.u_{\Lambda}(x,y,t)=u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t)+O\left(|t|^{-1/2}\right),\quad|t|\gg 1. (41)
  2. 2.

    If zero is a root of WΛ(z)W_{\Lambda}(z), i.e., d0d\neq 0 and 1-1, then in the inner region — the region that is within O(|t|1/3)O(|t|^{1/3}) of the wave center (x,y)=(12t,0)(x,y)=(12t,0), or (x12t)2+y2O(|t|1/3)\sqrt{(x-12t)^{2}+y^{2}}\leq O\left(|t|^{1/3}\right), lies d(d+1)/2d(d+1)/2 fundamental lumps u1(xx0,yy0,t)u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t), where u1(x,y,t)u_{1}(x,y,t) is given in Eq. (16),

    x0=(z0)(12t)1/3+O(1),y0=(z0)2(12t)1/3+O(1),x_{0}=\Re(z_{0})\hskip 1.42271pt(12t)^{1/3}+O(1),\quad y_{0}=\frac{\Im(z_{0})}{2}(12t)^{1/3}+O(1), (42)

    and z0z_{0} is each of the d(d+1)/2d(d+1)/2 simple roots of the Yablonskii–Vorob’ev polynomial Qd^(z)Q_{\hat{d}}(z), with d^\hat{d} defined as

    d^={d,whend0,|d|1,whend1.\hat{d}=\left\{\begin{array}[]{ll}d,&\mbox{when}\hskip 2.84544ptd\geq 0,\\ |d|-1,&\mbox{when}\hskip 2.84544ptd\leq-1.\end{array}\right. (43)

    Notice that d(d+1)/2=d^(d^+1)/2d(d+1)/2=\hat{d}(\hat{d}+1)/2. The absolute error of this fundamental-lump approximation is O(|t|1/3)O(|t|^{-1/3}) when z00z_{0}\neq 0 and O(|t|1)O(|t|^{-1}) when zero is a root of Qd^(z)Q_{\hat{d}}(z) and z0=0z_{0}=0. Expressed mathematically, when (x,y)(x,y) is in the neighborhood of each of these fundamental lumps, i.e., (x12tx0)2+(yy0)2=O(1)(x-12t-x_{0})^{2}+(y-y_{0})^{2}=O(1), with (x0,y0)(x_{0},y_{0}) given in (42), we have the following solution asymptotics for |t|1|t|\gg 1,

    uΛ(x,y,t)={u1(xx0,yy0,t)+O(|t|1/3),ifz00,u1(xx0,yy0,t)+O(|t|1),ifz0=0.u_{\Lambda}(x,y,t)=\left\{\begin{array}[]{ll}u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t)+O\left(|t|^{-1/3}\right),&\mbox{if}\hskip 1.9919ptz_{0}\neq 0,\\ u_{1}(x-x_{0},\hskip 1.13791pty-y_{0},t)+O\left(|t|^{-1}\right),&\mbox{if}\hskip 1.9919ptz_{0}=0.\end{array}\right. (44)
  3. 3.

    When (x,y)(x,y) is not in the neighborhood of any of the above fundamental lumps, uΛ(x,y,t)u_{\Lambda}(x,y,t) asymptotically approaches zero as |t||t|\to\infty.

Remark 6. In this theorem, we assumed all nonzero roots of WΛ(z)W_{\Lambda}(z) simple, which is true for all examples we tested, such as the two in Eqs. (32)-(33). In view of Conjecture 1 in the previous section, this assumption is expected to hold in all cases. If this conjecture is false, i.e., some nonzero roots of WΛ(z)W_{\Lambda}(z) are not simple, then this theorem for the outer region, i.e., Eqs. (40)-(41), would still hold, but only for nonzero simple roots z0z_{0} of WΛ(z)W_{\Lambda}(z).

Now, we explain what Theorem 3 says regarding solution patterns at large times when Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1). In this case, Theorem 3 indicates that, the whole wave field is generically split up into two regions featuring different patterns.

  1. 1.

    In the outer region — the region that is O(|t|1/2)O(|t|^{1/2}) away from the wave center (x,y)=(12t,0)(x,y)=(12t,0), the wave field at large time comprises NWN_{W} fundamental lumps, whose positions are given through the nonzero roots of the Wronskian-Hermite polynomial WΛ(z)W_{\Lambda}(z). Specifically, relative to the moving frame of xx-direction velocity 1212, positions (x0,y0)(x_{0},y_{0}) of these fundamental lumps, to the leading order of large time, are just a linear transformation of WΛ(z)W_{\Lambda}(z)’s nonzero-root structure. The reader is reminded from Sec. II.2.2 that when Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1), nonzero roots of WΛ(z)W_{\Lambda}(z) always exist, and their shape in the zz-plane is non-triangular. When tt is large negative, these fundamental-lump positions to the leading order are

    [x0y0]=(12|t|)1/2[10012][(z0)(z0)],\left[\begin{array}[]{c}x_{0}^{-}\\ y_{0}^{-}\end{array}\right]=(12|t|)^{1/2}\left[\begin{array}[]{cc}1&0\\ 0&\frac{1}{2}\end{array}\right]\left[\begin{array}[]{c}\Re(z_{0})\\ \Im(z_{0})\end{array}\right], (45)

    where z0z_{0} is any nonzero root of WΛ(z)W_{\Lambda}(z). However, when tt is large positive, these lump positions become

    [x0+y0+]=(12|t|)1/2[01120][(z0)(z0)].\left[\begin{array}[]{c}x_{0}^{+}\\ y_{0}^{+}\end{array}\right]=(12|t|)^{1/2}\left[\begin{array}[]{cc}0&-1\\ \frac{1}{2}&0\end{array}\right]\left[\begin{array}[]{c}\Re(z_{0})\\ \Im(z_{0})\end{array}\right]. (46)

    In the former case, the wave pattern formed by these fundamental lumps is simply a stretching of the Wronskian-Hermite nonzero-root structure along both horizontal and vertical directions. But in the latter case, on top of this stretching, the horizontal and vertical directions are also swapped. In both cases, the resulting wave patterns from transformations (45)-(46) are non-triangular since the root structure of WΛ(z)W_{\Lambda}(z) is non-triangular.

    From the above two transformations, we see that fundamental lumps at large negative time t-t and large positive time +t+t in the outer region are related as

    [x0+y0+]=[02120][x0y0].\left[\begin{array}[]{c}x_{0}^{+}\\ y_{0}^{+}\end{array}\right]=\left[\begin{array}[]{cc}0&-2\\ \frac{1}{2}&0\end{array}\right]\left[\begin{array}[]{c}x_{0}^{-}\\ y_{0}^{-}\end{array}\right]. (47)

    Thus, when time goes from large negative to large positive, outer-region lump patterns in the (x,y)(x,y) plane have swapped horizontal and vertical directions. In addition, stretching of different amounts has also occurred along the two directions. This swapping of horizontal and vertical directions is another type of dramatic pattern transformation, and it is very different from the triangular xx-direction reversal that occurs when Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1). For certain single-line patterns of fundamental lumps, a change from a vertical line to a horizontal line in the (x,y)(x,y) plane has been graphically reported in Chen2016 and analytically explained in Chang2018 . Here, we proved this fact for the general case, where patterns of fundamental lumps based on Wronskian-Hermite root structures can be arbitrary, not just lines (see the next section for examples).

    In this outer region, fundamental lumps at large time separate from each other in proportion to |t|1/2|t|^{1/2}. This is another big difference between the present solutions and those with Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1) in the previous subsection, where lumps separate in proportion to |t|1/3|t|^{1/3} instead.

  2. 2.

    In the inner region — the region that is within O(|t|1/3)O(|t|^{1/3}) of the wave center (x,y)=(12t,0)(x,y)=(12t,0), if d0d\neq 0 and 1-1, then the solution uΛ(x,y,t)u_{\Lambda}(x,y,t) at large time would comprise d(d+1)/2d(d+1)/2 fundamental lumps, whose positions are given through roots of the Yablonskii–Vorob’ev polynomial Qd^(z)Q_{\hat{d}}(z), with d^\hat{d} defined in Eq. (43). The reader is reminded that d^(d^+1)/2=d(d+1)/2\hat{d}(\hat{d}+1)/2=d(d+1)/2. Relative to the moving frame of xx-direction velocity 1212, positions (x0,y0)(x_{0},y_{0}) of these fundamental lumps, to the leading order of large time, are just a linear transformation of Qd^(z)Q_{\hat{d}}(z)’s root structure, i.e.,

    [x0y0]=(12t)1/3[10012][(z0)(z0)],\left[\begin{array}[]{c}x_{0}\\ y_{0}\end{array}\right]=(12t)^{1/3}\left[\begin{array}[]{cc}1&0\\ 0&\frac{1}{2}\end{array}\right]\left[\begin{array}[]{c}\Re(z_{0})\\ \Im(z_{0})\end{array}\right], (48)

    where z0z_{0} is each of the d(d+1)/2d(d+1)/2 simple roots of Qd^(z)Q_{\hat{d}}(z). This lump-position formula in the inner region is very similar to (38) of the Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1) case. Thus, the pattern of these d(d+1)/2d(d+1)/2 fundamental lumps in the inner region at large time is a simple stretching of Qd^(z)Q_{\hat{d}}(z)’s root structure, and the resulting pattern is triangular if d^>1\hat{d}>1. In addition, as time evolves from large negative to large positive, these triangular lump patterns would reverse direction along the xx-axis. Furthermore, fundamental lumps in this inner region separate from each other in proportion to |t|1/3|t|^{1/3} at large time, similar to the Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1) case in Theorem 2. If d^=0\hat{d}=0, i.e., d=0d=0 or 1-1, this inner region would be absent.

The above results reveal that, the pattern of the solution uΛ(x,y,t)u_{\Lambda}(x,y,t) for Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1) at large time is richer, with the outer region exhibiting the non-triangular shape of the stretched nonzero-root structure of the Wronskian-Hermite polynomial WΛ(z)W_{\Lambda}(z), and with the inner region exhibiting the triangular shape of the stretched root structure of the Yablonskii–Vorob’ev polynomial Qd^(z)Q_{\hat{d}}(z). As time changes from large negative to large positive, the outer pattern swaps horizontal and vertical directions, while the inner pattern reverses the horizontal direction. These different types of pattern transformations in the outer and inner regions of the same solution are fascinating. When Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1), the outer pattern is always present since WΛ(z)W_{\Lambda}(z) always has nonzero roots, but the inner pattern is present only when d0d\neq 0 and 1-1 and absent otherwise. When Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1), the outer pattern disappears, since WΛ(z)W_{\Lambda}(z) has only zero roots (see Lemma 1). In this special case, our results for the inner region in Theorem 3 are consistent with those in Theorem 2. However, Theorem 2 for this special case is stronger, since it shows that positions (x0,y0)(x_{0},y_{0}) of fundamental lumps now have no O(1)O(1) shifts [see Eq. (36)] — a more accurate prediction than Eq. (42) of Theorem 3 which shows O(1)O(1) shifts in general.

In the end, we note that in the earlier work Ablowitz2000 , it was reported that at large time, fundamental lumps in the higher-order lump complex separate from each other in proportion to |t|q|t|^{q}, where 13q12\frac{1}{3}\leq q\leq\frac{1}{2}. Our results in Theorems 2 and 3 indicate that this qq value can only be 1/31/3 or 1/21/2, nothing in between.

IV Comparison between true lump patterns and analytical predictions

In this section, we compare our analytical predictions of lump patterns with true solutions.

IV.1 Pattern transformation when Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1)

First, we do the comparison when Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1), where a triangular pattern of lumps at large time is predicted. To be specific, we take N=4N=4; so Λ=(1,3,5,7)\Lambda=(1,3,5,7). Root structure of the corresponding Yablonskii-Vorob’ev polynomial Q4(z)Q_{4}(z) has been displayed in Fig. 1. Using those roots and formulae (36), predicted solutions from Theorem 2 at large times t=10t=-10 and 1010 are plotted in Fig. 3.

Refer to caption
Figure 3: Predicted solutions uΛ(x,y,t)u_{\Lambda}(x,y,t) with Λ=(1,3,5,7)\Lambda=(1,3,5,7) at time values of t=10t=-10 (left) and t=10t=10 (right).

Now, we compare these predicted solutions with true ones. In the true solution uΛ(x,y,t)u_{\Lambda}(x,y,t), we select its internal parameters as a=(0,0,0,0,0,0,0)\textbf{\emph{a}}=(0,0,0,0,0,0,0). Then, evolutions of this true solution, at six time values of t=10,1, 0, 0.2, 1t=-10,\ -1,\ 0,\ 0.2,\ 1 and 1010, are plotted in Fig. 4. When comparing these true solutions at large times t=±10t=\pm 10 to those predicted in Fig. 3, they clearly match each other. First, the true solutions at t=±10t=\pm 10 indeed exhibit a triangular pattern, as the prediction says. Second, the triangular pattern at t=10t=10 is indeed a xx-direction reversal of the triangular pattern at t=10t=-10, relative to a frame moving with xx-direction velocity 1212. Thirdly, we have quantitatively compared the predicted and true solutions at t=±10t=\pm 10, and found them to match each other as well. For example, we have quantitatively compared the difference between the true peak location of individual lumps and its analytical prediction (36), similar to what we did in Figs. 5-6 of Ref. YangYang21a on the error analysis of rogue-pattern predictions. This comparison shows that this difference is indeed O(|t|1/3)O(|t|^{-1/3}) when z00z_{0}\neq 0 and O(|t|1)O(|t|^{-1}) when z0=0z_{0}=0, as our analytical formula (37) says. Thus, our asymptotic theory on patterns of higher-order lumps at large times for Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1) is fully confirmed.

In addition to large times, Fig. 4 also displays the true solution uΛ(x,y,t)u_{\Lambda}(x,y,t) at intermediate times, where our asymptotic theory does not apply. These intermediate panels shed light on how the dramatic xx-direction reversal of triangular patterns takes place as time changes from large negative to large positive. We see that in this solution, as time increases from 10-10 to 10, the triangle of fundamental lumps first approach each other and shrink in size, then coalesce at t=0t=0 and form a single lump of extreme height that is ten times that of original fundamental lumps, and then separate into a triangle of fundamental lumps again but with reversed xx-direction. This transformation process is fascinating.

Refer to caption
Refer to caption
Refer to caption
Figure 4: The true solution uΛ(x,y,t)u_{\Lambda}(x,y,t) with Λ=(1,3,5,7)\Lambda=(1,3,5,7) and a=(0,0,0,0,0,0,0)\textbf{\emph{a}}=(0,0,0,0,0,0,0) at various time values shown inside the panels.

How will this uΛ(x,y,t)u_{\Lambda}(x,y,t) solution evolve if its internal parameters a are different from (0,0,0,0,0,0,0)(0,0,0,0,0,0,0)? Theorem 2 tells us that in this case, the uΛ(x,y,t)u_{\Lambda}(x,y,t) solution would approach the same asymptotic state as that shown in Fig. 4 at large times. At intermediate times, however, this uΛ(x,y,t)u_{\Lambda}(x,y,t) solution could look very different from that in Fig. 4. For instance, by suitably choosing the a values, we can get uΛ(x,y,t)u_{\Lambda}(x,y,t) solutions whose graphs at t=0t=0 exhibit very different patterns such as a pentagon or a heptagon — a phenomenon that has been reported in Gaillard2018 . Thus, although these uΛ(x,y,t)u_{\Lambda}(x,y,t) solutions with different a values exhibit the same large-time triangular patterns, how this triangular pattern at large negative time transforms to its xx-reversed pattern at large positive time is a process that strongly depends on the choices of the internal a values.

IV.2 Pattern transformation when Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1)

Next, we perform the comparison when Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1), where the solution pattern at large time is determined by nonzero-root structure of the Wronskian-Hermit polynomial WΛ(z)W_{\Lambda}(z) in the outer region, and by root structure of the Yablonskii-Vorob’ev polynomial Qd^(z)Q_{\hat{d}}(z) in the inner region (if d^>0\hat{d}>0). Since this inner region can be present or absent depending on the dd value [see Eq. (26)], we will present two examples, one for each case.

Our first example is N=4N=4 and Λ=(2,3,4,5)\Lambda=(2,3,4,5). In this case, d=0d=0, and thus zero is not a root of WΛ(z)W_{\Lambda}(z) and the inner region is absent. Root structure of the corresponding Wronskian-Hermit polynomial has been displayed in Fig. 2 (the left panel). It was seen that this WΛ(z)W_{\Lambda}(z) admits eight simple nonzero roots which form a rectangle pattern. Using those roots and leading-order terms in formulae (40), predicted solutions from Theorem 3 at large times t=6t=-6 and 66 are plotted in Fig. 5. The predicted patterns contain eight fundamental lumps which also form a rectangular shape in the (x,y)(x,y) plane. At t=6t=-6, this lump pattern is just a stretching of the Wronskian-Hermit root structure. But at t=6t=6, this lump pattern has swapped its xx and yy directions and changed from its original xx-direction orientation to the new yy-direction orientation.

Refer to caption
Figure 5: Predicted solutions uΛ(x,y,t)u_{\Lambda}(x,y,t) with Λ=(2,3,4,5)\Lambda=(2,3,4,5) at time values of t=6t=-6 (left) and t=6t=6 (right).

To confirm these asymptotic predictions, we plot in Fig. 6 the corresponding true solution uΛ(x,y,t)u_{\Lambda}(x,y,t) at six time values of t=6,2,0.5,0,2t=-6,-2,-0.5,0,2 and 66. In this true solution, we have selected its internal parameters as a=(0,0,0,0,800)\textbf{\emph{a}}=(0,0,0,0,800). It is seen that at large times of t=±6t=\pm 6, the true solutions indeed comprise eight fundamental lumps forming a rectangular shape, and their orientations have changed from the xx-direction to the yy-direction, exactly as our asymptotic theory has predicted. In addition, quantitative comparisons between these true rectangular patterns and the predicted ones in Fig. 5 show good agreement. Thus, our asymptotic theory on patterns of higher-order lumps at large times is fully confirmed for Λ=(2,3,4,5)\Lambda=(2,3,4,5).

Refer to caption
Refer to caption
Refer to caption
Figure 6: True solutions uΛ(x,y,t)u_{\Lambda}(x,y,t) with Λ=(2,3,4,5)\Lambda=(2,3,4,5) and a=(0,0,0,0,800)\textbf{\emph{a}}=(0,0,0,0,800), at various times whose values are shown inside the panels.

By inspecting Fig. 6, we can also see how this dramatic rectangular-pattern reorientation takes place as time increases. First, these eight fundamental lumps of rectangular shape with xx-direction orientation get closer to each other and rearrange their shapes. At t=0t=0, the solution has evolved into a pentagon of five fundamental lumps surrounding a higher-peak lump near the center. Afterwards, this pentagon structure further adjusts its shape in significant ways, until eight new fundamental lumps emerge as a rectangular with yy-direction orientation in the end. Again, this transformation process is amazing.

Our second example is N=5N=5 and Λ=(3,4,5,7,9)\Lambda=(3,4,5,7,9). In this case, d=3d=3, and thus zero is a root of multiplicity six in WΛ(z)W_{\Lambda}(z), and the inner region is present. Root structure of the corresponding Wronskian-Hermit polynomial has been displayed in Fig. 2 (the right panel). It is seen that this WΛ(z)W_{\Lambda}(z) admits 12 simple nonzero roots which form a quasi-rectangular shape, plus the zero root of multiplicity six at the center of the quasi-rectangle. Using those roots and leading-order terms in formulae (40) and (42), predicted solutions from Theorem 3 at large times t=10t=-10 and 1010 are plotted in Fig. 7. The predicted patterns contain 12 fundamental lumps which also form a quasi-rectangular pattern in the outer region of the (x,y)(x,y) plane, plus six fundamental lumps which form a triangle in the inner region. At t=10t=-10, the outer lump pattern is a stretching of the Wronskian-Hermit polynomial WΛ(z)W_{\Lambda}(z)’s nonzero-root structure, while the inner lump pattern is a stretching of the Yablonskii–Vorob’ev polynomial Q3(z)Q_{3}(z)’s root structure. At t=10t=10, however, the predicted outer lump pattern has rotated by 9090^{\circ} from its t=10t=-10 state [plus additional (x,y)(x,y)-direction stretching], while the predicted inner triangular lump pattern has reversed its direction along the xx-axis.

Refer to caption
Figure 7: Predicted solutions uΛ(x,y,t)u_{\Lambda}(x,y,t) with Λ=(3,4,5,7,9)\Lambda=(3,4,5,7,9) at time values of t=10t=-10 (left) and t=10t=10 (right).

To confirm these asymptotic predictions, we plot in Fig. 8 the corresponding true solution uΛ(x,y,t)u_{\Lambda}(x,y,t) at six time values of t=10,2,0.2,0,2t=-10,-2,-0.2,0,2 and 1010. In this true solution, we have selected all-zero internal parameters of a=(0,0,0,0,0,0,0,0,0)\textbf{\emph{a}}=(0,0,0,0,0,0,0,0,0). It is seen that at large times of t=±10t=\pm 10, the true solutions closely match our predictions in the previous figure. Specifically, the true solutions at these large times also split into outer and inner regions, with outer quasi-rectangular patterns and inner triangular patterns closely resembling our predicted ones in Fig. 7. Quantitative comparisons between these true patterns and predicted ones show good agreement as well. Thus, our asymptotic theory on higher-order lump patterns at large times is fully confirmed for Λ=(3,4,5,7,9)\Lambda=(3,4,5,7,9).

Refer to caption
Refer to caption
Refer to caption
Figure 8: True solutions uΛ(x,y,t)u_{\Lambda}(x,y,t) with Λ=(3,4,5,7,9)\Lambda=(3,4,5,7,9) and a=(0,0,0,0,0,0,0,0,0)\textbf{\emph{a}}=(0,0,0,0,0,0,0,0,0), at various times whose values are shown inside the panels.

True solution graphs at intermediate time values in Fig. 8 reveal how these striking pattern transformations in outer and inner regions take place. It is seen that all fundamental lumps in the inner and outer regions at large negative time first move toward each other. Then they merge and coalesce at t0t\approx 0. Afterwards, all these fundamental lumps re-emerge and move away from each other, but not returning to their pre-merging state. Instead, the quasi-rectangular outer lumps have swapped their xx and yy directions, and the triangular inner lumps have reversed the xx-direction. These pattern transformations are visually miraculous and mysterious. But due to our Theorem 3, they can now be completely understood from a mathematical point of view.

V Proofs of the two theorems

Now, we prove our two theorems stated in Sec. 3. The reader is reminded that in these proofs, p=1p=1 and ai\textbf{\emph{a}}_{i} are chosen as (35) in the higher-order lump solutions of Theorem 1, for reasons which have been explained earlier in the paper. Thus, solution expressions in Theorem 1 can be simplified. Notably, the constant factor in Eq. (9) simplifies to 1/4ν1/4^{\nu}, and the s vector is real with sodd=0s_{odd}=0 (see the beginning of Sec. 3).

V.1 Proof of Theorem 2

In this case, Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1). First, we rewrite the determinant (8) as a larger 3N×3N3N\times 3N determinant OhtaJY2012 ; YangYang21a

σ=|ON×NΦN×2NΨ2N×NI2N×2N|,\sigma=\left|\begin{array}[]{cc}\textbf{O}_{N\times N}&\Phi_{N\times 2N}\\ -\Psi_{2N\times N}&\textbf{I}_{2N\times 2N}\end{array}\right|, (49)

where

Φi,j=2(j1)S2ij[x++(j1)s+a],Ψi,j=2(i1)S2ji[(x+)+(i1)s+a],\Phi_{i,j}=2^{-(j-1)}S_{2i-j}\left[\textbf{\emph{x}}^{+}+(j-1)\textbf{\emph{s}}+\textbf{\emph{a}}\right],\quad\Psi_{i,j}=2^{-(i-1)}S_{2j-i}\left[(\textbf{\emph{x}}^{+})^{*}+(i-1)\textbf{\emph{s}}+\textbf{\emph{a}}^{*}\right], (50)

Sj0S_{j}\equiv 0 for j<0j<0, and vectors x+\textbf{\emph{x}}^{+} and s are given in Eqs. (10) and (34). This determinant can be further simplified. Indeed, using the technique outlined in Appendix A of Ref. YangYang21a , we can eliminate all xeven+x_{even}^{+} and aevena_{even} terms from the vectors x+\textbf{\emph{x}}^{+} and a, and reduce the above matrix element formulae to

Φi,j=2(j1)S2ij[x^++(j1)s+a^],Ψi,j=2(i1)S2ji[(x^+)+(i1)s+a^],\Phi_{i,j}=2^{-(j-1)}S_{2i-j}\left[\hat{\textbf{\emph{x}}}^{+}+(j-1)\textbf{\emph{s}}+\hat{\textbf{\emph{a}}}\right],\quad\Psi_{i,j}=2^{-(i-1)}S_{2j-i}\left[(\hat{\textbf{\emph{x}}}^{+})^{*}+(i-1)\textbf{\emph{s}}+\hat{\textbf{\emph{a}}}^{*}\right], (51)

where

x^+(x1+,0,x3+,0,x5+,0,),a^(0,0,a3,0,a5,0,).\hat{\textbf{\emph{x}}}^{+}\equiv\left(x_{1}^{+},0,x_{3}^{+},0,x_{5}^{+},0,\cdots\right),\quad\hat{\textbf{\emph{a}}}\equiv\left(0,0,a_{3},0,a_{5},0,\cdots\right). (52)

The elimination of the solution’s dependence on x2+x_{2}^{+} is a key feature of the index vector Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1), and this feature is responsible for the distinctive pattern behaviors described in Theorem 2.

Now, we analyze the large-time asymptotics of the above determinant σ\sigma. For this purpose, we introduce a moving xx-frame coordinate

x^x12t.\hat{x}\equiv x-12t. (53)

Then, the elements xk+x_{k}^{+} in Eq. (10) become

x1+=x^+2iy,xk+=1k!x^+2kk!iy+Tk,x_{1}^{+}=\hat{x}+2\textrm{i}y,\quad x_{k}^{+}=\frac{1}{k!}\hat{x}+\frac{2^{k}}{k!}\textrm{i}y+T_{k}, (54)

where Tk12(13k1)t/k!T_{k}\equiv 12(1-3^{k-1})t/k!. In particular,

T2=12t,T3=16t.T_{2}=-12t,\quad T_{3}=-16t. (55)

In this moving xx-frame, when |t||t| is large and x^2+y2=O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/3}), we have the leading-order asymptotics for Sk(x^++νs+a^)S_{k}\left(\hat{\textbf{\emph{x}}}^{+}+\nu\textbf{\emph{s}}+\hat{\textbf{\emph{a}}}\right) as

Sk(x^++νs+a^)Sk(v),|t|1,S_{k}\left(\hat{\textbf{\emph{x}}}^{+}+\nu\textbf{\emph{s}}+\hat{\textbf{\emph{a}}}\right)\sim S_{k}(\textbf{v}),\quad\quad|t|\gg 1, (56)

where

v=(x1+,0,T3,0,0,0,).\textbf{v}=\left(x_{1}^{+},\hskip 1.13791pt0,\hskip 1.13791ptT_{3},\hskip 1.13791pt0,\hskip 1.13791pt0,\hskip 1.13791pt0,\cdots\right). (57)

By comparing the definition of Schur polynomials Sk(v)S_{k}(\textbf{v}) to the definition of pk(z)p_{k}(z) polynomials in Eq. (17), we see that

Sk(v)=(3T3/4)k/3pk(z),S_{k}(\textbf{v})=\left(-3T_{3}/4\right)^{k/3}p_{k}(z), (58)

where

z=(3T3/4)1/3x1+=(3T3/4)1/3(x^+2iy).z=(-3T_{3}/4)^{-1/3}x_{1}^{+}=(-3T_{3}/4)^{-1/3}\left(\hat{x}+2\textrm{i}y\right). (59)

Using these formulae and the Laplace expansion of the 3N×3N3N\times 3N determinant (49)

σ=0ν1<ν2<<νN2N1det1i,jN[12νjS2i1νj(x^++νjs+a^)]×det1i,jN[12νjS2i1νj[(x^+)+νjs+a^]],\displaystyle\sigma=\sum_{0\leq\nu_{1}<\nu_{2}<\cdots<\nu_{N}\leq 2N-1}\det_{1\leq i,j\leq N}\left[\frac{1}{2^{\nu_{j}}}S_{2i-1-\nu_{j}}(\hat{\textbf{\emph{x}}}^{+}+\nu_{j}\textbf{\emph{s}}+\hat{\textbf{\emph{a}}})\right]\times\det_{1\leq i,j\leq N}\left[\frac{1}{2^{\nu_{j}}}S_{2i-1-\nu_{j}}[(\hat{\textbf{\emph{x}}}^{+})^{*}+\nu_{j}\textbf{\emph{s}}+\hat{\textbf{\emph{a}}}^{*}]\right], (60)

together with the fact that the highest order term of |t||t| in this σ\sigma comes from the index choices of νj=j1\nu_{j}=j-1, we can readily show that the highest tt-power term of σ\sigma is

σ|α0|2|3T3/4|N(N+1)3|QN(z)|2,|t|1,\sigma\sim|\alpha_{0}|^{2}\hskip 1.42271pt\left|3T_{3}/4\right|^{\frac{N(N+1)}{3}}\left|Q_{N}(z)\right|^{2},\quad\quad|t|\gg 1, (61)

where α0=2N(N1)/2cN1\alpha_{0}=2^{-N(N-1)/2}c_{N}^{-1}. Inserting this leading-order term of σ\sigma into Eq. (7), we see that the solution uΛ(x,y,t)u_{\Lambda}(x,y,t) approaches zero when |t||t|\to\infty, except at or near (x^,y)(\hat{x},y) locations (x0,y0)\left(x_{0},y_{0}\right), i.e., at or near (x,y)(x,y) locations (12t+x0,y0)\left(12t+x_{0},y_{0}\right), where

z0=(3T3/4)1/3(x0+2iy0)z_{0}=(-3T_{3}/4)^{-1/3}\left(x_{0}+2\textrm{i}y_{0}\right) (62)

is a root of the polynomial QN(z)Q_{N}(z). Solving this equation, we get the (x0,y0)(x_{0},y_{0}) locations given by Eq. (36) in Theorem 2. Due to our requirement of x^2+y2=O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/3}), z0z_{0} in the above equation should be nonzero.

In order to derive the solution behavior near this (x,y)=(12t+x0,y0)(x,y)=\left(12t+x_{0},y_{0}\right) location, we need to perform a more refined asymptotic analysis and calculate the next-order terms in tt, since the leading-order term in Eq. (61) vanishes at this point. Recalling s1=0s_{1}=0, this refined analysis is very similar to that we did for rogue waves in the NLS equation YangYang21a . For z00z_{0}\neq 0 in the (x0,y0)(x_{0},y_{0}) formula (36), i.e., if the (12t+x0,y0)\left(12t+x_{0},y_{0}\right) location is O(|t|1/3)O(|t|^{1/3}) away from the wave center (12t,0)(12t,0), then in the O(1)O(1) neighborhood of (12t+x0,y0)\left(12t+x_{0},y_{0}\right), i.e., when (x12tx0)2+(yy0)2=O(1)(x-12t-x_{0})^{2}+(y-y_{0})^{2}=O(1), we have an asymptotics more refined than (56), which is

Sk(x^++νs+a^)=Sk(v)[1+O(|t|2/3)].S_{k}\left(\hat{\textbf{\emph{x}}}^{+}+\nu\textbf{\emph{s}}+\hat{\textbf{\emph{a}}}\right)=S_{k}(\textbf{v})\left[1+O(|t|^{-2/3})\right]. (63)

This O(|t|2/3)O(|t|^{-2/3}) relative error is due to our omission of x^/6+4iy/3\hat{x}/6+4\textrm{i}y/3 relative to T3T_{3} in x3+x_{3}^{+}, and omission of x5+x_{5}^{+} relative to x3+x_{3}^{+}. Using this refined asymptotics and repeating the same steps as in YangYang21a , we find that

σ(x,y,t)=|α0|2|QN(z0)|2|3T3/4|N(N+1)23[(x12tx0)2+4(yy0)2+14][1+O(|t|1/3)],\displaystyle\sigma(x,y,t)=\left|\alpha_{0}\right|^{2}\hskip 1.70709pt\left|Q_{N}^{\prime}(z_{0})\right|^{2}|3T_{3}/4|^{\frac{N(N+1)-2}{3}}\left[\left(x-12t-x_{0}\right)^{2}+4(y-y_{0})^{2}+\frac{1}{4}\right]\left[1+O\left(|t|^{-1/3}\right)\right], (64)

where α0\alpha_{0} is given below Eq. (61). For Yablonskii-Vorob’ev polynomials QN(z)Q_{N}(z), all roots are simple. Thus, QN(z0)0Q_{N}^{\prime}(z_{0})\neq 0.

In the O(1)O(1) neighborhood of the wave center (12t,0)\left(12t,0\right), where (x12t)2+y2=O(1)(x-12t)^{2}+y^{2}=O(1), we need to perform a separate asymptotic analysis, because the earlier SkS_{k} asymptotics (56) and (63) do not hold in this region. In this case, due to Eq. (54), when we lump T2k+1T_{2k+1} and a2k+1a_{2k+1} together in Eq. (51) and recall T2k+1T_{2k+1} is proportional to tt, the large-time analysis of the present σ\sigma determinant (49) is almost identical to that in Appendix C of Ref. YangYang21a for the analysis of NLS rogue patterns when its internal parameters (a3,a5,)(a_{3},a_{5},\cdots) are all large and of the same order. Repeating that analysis, we find that if zero is a root of the Yablonskii-Vorob’ev polynomial QN(z)Q_{N}(z), i.e., N1mod3N\equiv 1\hskip 2.84544pt\mbox{mod}\hskip 2.84544pt3, then

σ(x,y,t)=β0|t|N(N+1)23[(x12t)2+4y2+14][1+O(|t|1)],\displaystyle\sigma(x,y,t)=\beta_{0}\hskip 1.42271pt|t|^{\frac{N(N+1)-2}{3}}\left[\left(x-12t\right)^{2}+4y^{2}+\frac{1}{4}\right]\left[1+O\left(|t|^{-1}\right)\right], (65)

where β0\beta_{0} is a NN-dependent positive constant. If zero is not a root of QN(z)Q_{N}(z), then σ(x,y,t)β0|t|(N+2)(N1)3\sigma(x,y,t)\sim\beta_{0}\hskip 1.42271pt|t|^{\frac{(N+2)(N-1)}{3}}.

Substituting the above two σ\sigma asymptotics (64)-(65) into the solution expression (7) and performing a little simplification, we then get the asymptotics (37). Theorem 2 is then proved.

V.2 Proof of Theorem 3

In this case, Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1). We first rewrite the determinant σ\sigma in (8) as a larger (N+nN+1)×(N+nN+1)(N+n_{N}+1)\times(N+n_{N}+1) determinant

σ=|ON×NΦN×(nN+1)Ψ(nN+1)×NI(nN+1)×(nN+1)|,\sigma=\left|\begin{array}[]{cc}\textbf{O}_{N\times N}&\Phi_{N\times(n_{N}+1)}\\ -\Psi_{(n_{N}+1)\times N}&\textbf{I}_{(n_{N}+1)\times(n_{N}+1)}\end{array}\right|, (66)

where

Φi,j=2(j1)Sni+1j[x++(j1)s+a],Ψi,j=2(i1)Snj+1i[(x+)+(i1)s+a],\Phi_{i,j}=2^{-(j-1)}S_{n_{i}+1-j}\left[\textbf{\emph{x}}^{+}+(j-1)\textbf{\emph{s}}+\textbf{\emph{a}}\right],\quad\Psi_{i,j}=2^{-(i-1)}S_{n_{j}+1-i}\left[(\textbf{\emph{x}}^{+})^{*}+(i-1)\textbf{\emph{s}}+\textbf{\emph{a}}^{*}\right], (67)

and vectors x+\textbf{\emph{x}}^{+} and s are given in Eqs. (10) and (34). Unlike the previous case, we cannot eliminate x2+x_{2}^{+} from this solution now. Our large-time asymptotics of this determinant proceeds as follows.

V.2.1 Proof for the outer region

First, we prove the asymptotics (40)-(41) for the outer region. In this region, x^2+y2=O(|t|1/2)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/2}). Thus, we have the leading-order asymptotics for Sk(x++νs+a)S_{k}\left(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}}\right) as

Sk(x++νs+a)Sk(w),|t|1,S_{k}\left(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}}\right)\sim S_{k}(\textbf{w}),\quad\quad|t|\gg 1, (68)

where

w=(x1+,T2,0,0,0,),\textbf{w}=\left(x_{1}^{+},\hskip 1.13791ptT_{2},\hskip 1.13791pt0,\hskip 1.13791pt0,\hskip 1.13791pt0,\cdots\right), (69)

and T2T_{2} is as given in Eq. (55). By comparing the definition of Schur polynomials Sk(w)S_{k}(\textbf{w}) to the definition of qk(z)q_{k}(z) polynomials in Eq. (23), we see that

Sk(w)=T2k/2qk(z),S_{k}(\textbf{w})=T_{2}^{k/2}q_{k}(z), (70)

where

z=T21/2x1+=T21/2(x^+2iy).z=T_{2}^{-1/2}x_{1}^{+}=T_{2}^{-1/2}\left(\hat{x}+2\textrm{i}y\right). (71)

Using these formulae and the Laplace expansion of the determinant (66) for σ\sigma, we can readily show that the highest tt-power term of σ\sigma is

σ|μ0|2|T2|ρ|WΛ(z)|2,|t|1,\sigma\sim|\mu_{0}|^{2}\hskip 1.42271pt|T_{2}|^{\rho}\left|W_{\Lambda}(z)\right|^{2},\quad\quad|t|\gg 1, (72)

where ρ\rho is given in Eq. (13), and μ0=2N(N1)/2\mu_{0}=2^{-N(N-1)/2}. Inserting this leading-order term of σ\sigma into Eq. (7), we see that the solution uΛ(x,y,t)u_{\Lambda}(x,y,t) approaches zero when |t||t|\to\infty, except at or near (x^,y)(\hat{x},y) locations (x^0,y^0)\left(\hat{x}_{0},\hat{y}_{0}\right), i.e., at or near (x,y)(x,y) locations (12t+x^0,y^0)\left(12t+\hat{x}_{0},\hat{y}_{0}\right), where

z0=T21/2(x^0+2iy^0)z_{0}=T_{2}^{-1/2}\left(\hat{x}_{0}+2\textrm{i}\hat{y}_{0}\right) (73)

is a root of the Wronskian-Hermit polynomial WΛ(z)W_{\Lambda}(z). Solving this equation, we get

x^0=[z0T21/2],y^0=[z0T21/2]2,\hat{x}_{0}=\Re\left[z_{0}T_{2}^{1/2}\right],\quad\hat{y}_{0}=\frac{\Im\left[z_{0}T_{2}^{1/2}\right]}{2}, (74)

which are the leading-order terms of (x0,y0)(x_{0},y_{0}) in Eq. (40) of Theorem 3. Due to our requirement of x^2+y2=O(|t|1/2)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/2}), z0z_{0} in the above equation should be nonzero.

To derive the solution behavior near this (x,y)=(12t+x^0,y^0)(x,y)=\left(12t+\hat{x}_{0},\hat{y}_{0}\right) location, we perform a more refined asymptotic analysis. Our starting point is a more accurate asymptotics for Sk(x++νs+a)S_{k}\left(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}}\right),

Sk(x++νs+a)=Sk(w^)[1+O(|t|1)],|t|1,S_{k}(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}})=S_{k}(\hat{\textbf{w}})\left[1+O\left(|t|^{-1}\right)\right],\quad\quad|t|\gg 1, (75)

where

w^=(x1+,x2+,T3,0,0,0,)=w+(0,x^2+,T3,0,0,0,),\hat{\textbf{w}}=\left(x_{1}^{+},\hskip 1.13791ptx_{2}^{+},\hskip 1.13791ptT_{3},\hskip 1.13791pt0,\hskip 1.13791pt0,\hskip 1.13791pt0,\cdots\right)=\textbf{w}+\left(0,\hat{x}_{2}^{+},T_{3},\hskip 1.13791pt0,\hskip 1.13791pt0,\hskip 1.13791pt0,\cdots\right), (76)

w is given in (69), and

x^2+12x^+2iy.\hat{x}_{2}^{+}\equiv\frac{1}{2}\hat{x}+2\textrm{i}y. (77)

The asymptotics (75) holds since a1=s1=0a_{1}=s_{1}=0. From the definition (6) of Schur polynomials and the above equation, we can relate Sk(w^)S_{k}(\hat{\textbf{w}}) and Sk(w)S_{k}(\textbf{w}) as

Sk(w^)=j=0kbjSkj(w),S_{k}(\hat{\textbf{w}})=\sum_{j=0}^{k}b_{j}S_{k-j}(\textbf{w}), (78)

where bjb_{j} are the coefficients in the expansion

ex^2+ϵ2+T3ϵ3=j=0bjϵj.e^{\hat{x}_{2}^{+}\epsilon^{2}+T_{3}\hskip 0.56917pt\epsilon^{3}}=\sum_{j=0}^{\infty}b_{j}\epsilon^{j}. (79)

Notice that b0=1b_{0}=1, b1=0b_{1}=0, b2=x^2+b_{2}=\hat{x}_{2}^{+}, and b3=T3b_{3}=T_{3}. In addition, (x^,y)=O(|t|1/2)(\hat{x},y)=O(|t|^{1/2}) from Eq. (74), and Sk(w)=O(|t|k/2)S_{k}(\textbf{w})=O(|t|^{k/2}) in view of Eq. (70). Utilizing these relations, we find that

Sk(x++νs+a)=[Sk(w)+x^2+Sk2(w)+T3Sk3(w)][1+O(|t|1)],|t|1.S_{k}(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}})=\left[S_{k}(\textbf{w})+\hat{x}_{2}^{+}S_{k-2}(\textbf{w})+T_{3}\hskip 0.56917ptS_{k-3}(\textbf{w})\right]\left[1+O\left(|t|^{-1}\right)\right],\quad\quad|t|\gg 1. (80)

With this formula (80), we can now determine the asymptotic expression of σ\sigma in Eq. (66) in the neighborhood of (x,y)=(12t+x^0,y^0)(x,y)=\left(12t+\hat{x}_{0},\hat{y}_{0}\right) at large tt. The Laplace expansion of this determinant is very similar to Eq. (60) of the previous subsection. Using this Laplace expansion and similar techniques as in Refs. YangYang21a ; YangYang21b , we can readily find that

σ(x,y,t)=|μ0|2|WΛ(z0)|2|T2|ρ1(|(x12tx^0)+2i(yy^0)+Δ|2+14)[1+O(|t|1/2)],\displaystyle\sigma(x,y,t)=\left|\mu_{0}\right|^{2}\hskip 1.70709pt\left|W_{\Lambda}^{\prime}(z_{0})\right|^{2}|T_{2}|^{\rho-1}\left(\left|(x-12t-\hat{x}_{0})+2\textrm{i}(y-\hat{y}_{0})+\Delta\right|^{2}+\frac{1}{4}\right)\left[1+O\left(|t|^{-1/2}\right)\right], (81)

where μ0\mu_{0} is given below Eq. (72), and Δ=Δ(Λ,z0)\Delta=\Delta(\Lambda,z_{0}) is an O(1)O(1) complex constant given by

Δ=1WΛ(z0){λj=1Ndet1iN[qni,,qni(j2),qni(j1)2,qnij,,qni(N1)]z=z0\displaystyle\Delta=\frac{1}{W_{\Lambda}^{\prime}(z_{0})}\left\{\lambda\sum_{j=1}^{N}\det_{1\leq i\leq N}\left[q_{n_{i}},\cdots,q_{n_{i}-(j-2)},\hskip 1.42271ptq_{n_{i}-(j-1)-2},\hskip 1.42271ptq_{n_{i}-j},\cdots,q_{n_{i}-(N-1)}\right]_{z=z_{0}}\right.
+43j=1Ndet1iN[qni,,qni(j2),qni(j1)3,qnij,,qni(N1)]z=z0},\displaystyle\hskip 51.21504pt\left.+\frac{4}{3}\sum_{j=1}^{N}\det_{1\leq i\leq N}\left[q_{n_{i}},\cdots,q_{n_{i}-(j-2)},\hskip 1.42271ptq_{n_{i}-(j-1)-3},\hskip 1.42271ptq_{n_{i}-j},\cdots,q_{n_{i}-(N-1)}\right]_{z=z_{0}}\right\}, (82)

and

λ={12(z0)+i(z0),whent<0,(z0)+12i(z0),whent>0.\lambda=\left\{\begin{array}[]{ll}\frac{1}{2}\Re(z_{0})+\textrm{i}\Im(z_{0}),&\mbox{when}\hskip 2.84544ptt<0,\\ \Re(z_{0})+\frac{1}{2}\textrm{i}\Im(z_{0}),&\mbox{when}\hskip 2.84544ptt>0.\end{array}\right. (83)

The former determinant in Eq. (82) is the Wronskian-Hermit determinant in Eq. (25) but with the jj-th column {qni(j1)}\{q_{n_{i}-(j-1)}\} replaced by {qni(j1)2}\{q_{n_{i}-(j-1)-2}\}, i.e., reducing the subindex value of this column by two, while the latter determinant in (82) is the Wronskian-Hermit determinant (25) with the jj-th column replaced by {qni(j1)3}\{q_{n_{i}-(j-1)-3}\}, i.e., reducing its subindex value by three.

The complex constant Δ\Delta in Eq. (81) can be absorbed into (x^0,y^0)(\hat{x}_{0},\hat{y}_{0}). After this absorption and rearranging terms, Eq. (81) becomes

σ(x,y,t)=|μ0|2|WΛ(z0)|2|T2|ρ1[(x12tx0)2+4(yy0)2+14][1+O(|t|1/2)],\displaystyle\sigma(x,y,t)=\left|\mu_{0}\right|^{2}\hskip 1.70709pt\left|W_{\Lambda}^{\prime}(z_{0})\right|^{2}|T_{2}|^{\rho-1}\left[\left(x-12t-x_{0}\right)^{2}+4(y-y_{0})^{2}+\frac{1}{4}\right]\left[1+O\left(|t|^{-1/2}\right)\right], (84)

where

x0=[z0(12t)1/2](Δ),y0=[z0(12t)1/2]2(Δ)2.x_{0}=\Re\left[z_{0}(-12t)^{1/2}\right]-\Re(\Delta),\quad y_{0}=\frac{\Im\left[z_{0}(-12t)^{1/2}\right]}{2}-\frac{\Im(\Delta)}{2}. (85)

These (x0,y0)(x_{0},y_{0}) formulae contain the explicit O(1)O(1) corrections to the leading O(|t|1/2)O(|t|^{1/2}) terms, and are thus more complete than Eq. (40) in Theorem 3. Substituting the above σ\sigma asymptotics (84) into Eq. (7), the asymptotics (40)-(41) for the outer region of Theorem 3 are then proved.

V.2.2 Proof for the inner region

In the inner region, where x^2+y2O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}\leq O(|t|^{1/3}), a separate asymptotic analysis is needed, because the previous SkS_{k} asymptotics (68) and (75) do not hold. In this inner region, our analysis needs to split into two cases, depending on whether x^2+y2\sqrt{\hat{x}^{2}+y^{2}} is O(|t|1/3)O(|t|^{1/3}) or O(1)O(1).

When x^2+y2=O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/3}), it is easy to see from the Laplace expansion of the σ\sigma determinant (66) that, at large |t||t|, the highest tt-power term of σ\sigma comes from the index choices of νj=j1\nu_{j}=j-1, i.e.,

σ|det1i,jN[12j1Sni+1j(x++(j1)s+a)]|2,|t|1.\displaystyle\sigma\sim\left|\det_{1\leq i,j\leq N}\left[\frac{1}{2^{j-1}}S_{n_{i}+1-j}(\textbf{\emph{x}}^{+}+(j-1)\textbf{\emph{s}}+\textbf{\emph{a}})\right]\right|^{2},\qquad|t|\gg 1. (86)

For the determinant in the above equation, we reorganize its rows by grouping odd-nin_{i} rows together (in ascending order of nin_{i}), followed by even-nin_{i} rows (also in ascending order of nin_{i}). We also rewrite Sk(x++νs+a)S_{k}(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}}) as

Sk(x++νs+a)=j=0[k/2]T2jj!Sk2j(h+νs+a),S_{k}(\textbf{\emph{x}}^{+}+\nu\textbf{\emph{s}}+\textbf{\emph{a}})=\sum_{j=0}^{\left[k/2\right]}\frac{T_{2}^{j}}{j!}S_{k-2j}(\textbf{h}+\nu\textbf{\emph{s}}+\textbf{\emph{a}}), (87)

where

hx+(0,T2,0,0,)=(x1+,x^2+,x3+,x4+,),\textbf{h}\equiv\textbf{\emph{x}}^{+}-(0,T_{2},0,0,\cdots)=(x_{1}^{+},\hat{x}_{2}^{+},x_{3}^{+},x_{4}^{+},\cdots), (88)

and [a][a] represents the largest integer less than or equal to aa. In addition, we notice that when x^2+y2=O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/3}),

Sk(h+νs+a)Sk(h^),S_{k}(\textbf{h}+\nu\textbf{\emph{s}}+\textbf{\emph{a}})\sim S_{k}(\hat{\textbf{h}}), (89)

where

h^=(x1+,0,T3,0,0,),\hat{\textbf{h}}=(x_{1}^{+},0,T_{3},0,0,\cdots), (90)

and

Sk(h^)=(3T3/4)k/3pk(z),S_{k}(\hat{\textbf{h}})=\left(-3T_{3}/4\right)^{k/3}p_{k}(z), (91)

where zz is as given in Eq. (59). Inserting these formulae into the above reorganized determinant and performing row operations to eliminate certain high powers of T2T_{2} in lower rows of the odd-nin_{i}-index group as well as the even-nin_{i}-index group, we find that the highest tt-power term of σ\sigma from Eq. (86) is

σγ0|T2|NW|3T3/4|d^(d^+1)3|H(z)|2,\sigma\sim\gamma_{0}\hskip 1.42271pt|T_{2}|^{N_{W}}|3T_{3}/4|^{\frac{\hat{d}(\hat{d}+1)}{3}}\left|H(z)\right|^{2}, (92)

where γ0\gamma_{0} is a certain positive constant, NWN_{W} is given in Eqs. (27)-(28), d^\hat{d} is defined in Eq. (43), H(z)H(z) is the determinant

H(z)=det[p1(z)p0(z)p1(z)p3(z)p2(z)p1(z)p2kodd1(z)p2kodd2(z)p2kodd3(z)p0(z)p1(z)p2(z)p2(z)p1(z)p0(z)p2keven2(z)p2keven3(z)p2keven4(z)],H(z)=\det\left[\begin{array}[]{cccc}p_{1}(z)&p_{0}(z)&p_{-1}(z)&\cdots\\ p_{3}(z)&p_{2}(z)&p_{1}(z)&\cdots\\ \vdots&\vdots&\vdots&\vdots\\ p_{2k_{odd}-1}(z)&p_{2k_{odd}-2}(z)&p_{2k_{odd}-3}(z)&\cdots\\ p_{0}(z)&p_{-1}(z)&p_{-2}(z)&\cdots\\ p_{2}(z)&p_{1}(z)&p_{0}(z)&\cdots\\ \vdots&\vdots&\vdots&\vdots\\ p_{2k_{even}-2}(z)&p_{2k_{even}-3}(z)&p_{2k_{even}-4}(z)&\cdots\end{array}\right], (93)

and koddk_{odd}, kevenk_{even} are the numbers of odd and even elements in the index vector (n1,n2,,nN)(n_{1},n_{2},\dots,n_{N}) respectively. Clearly, this HH determinant can be reduced to

H(z)=Qd^(z).H(z)=Q_{\hat{d}}(z). (94)

Thus, when x^2+y2=O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/3}),

σγ0|T2|NW|3T3/4|d^(d^+1)3|Qd^(z)|2.\sigma\sim\gamma_{0}\hskip 1.42271pt|T_{2}|^{N_{W}}|3T_{3}/4|^{\frac{\hat{d}(\hat{d}+1)}{3}}\left|Q_{\hat{d}}(z)\right|^{2}. (95)

In view of Eq. (7), this asymptotics shows that the solution uΛ(x,y,t)u_{\Lambda}(x,y,t) is asymptotically zero in this region, except when d^>0\hat{d}>0 (i.e., zero is a root of the Wronskian-Hermit polynomial WΛ(z)W_{\Lambda}(z)), and when (x,y)(x,y) is at or near the location (12t+x^0,y^0)\left(12t+\hat{x}_{0},\hat{y}_{0}\right), where

z0=(3T3/4)1/3(x^0+2iy^0)z_{0}=(-3T_{3}/4)^{-1/3}\left(\hat{x}_{0}+2\textrm{i}\hat{y}_{0}\right) (96)

is a root of the Yablonskii–Vorob’ev polynomial Qd^(z)Q_{\hat{d}}(z). Solving this equation, we get (x^0,y^0)(\hat{x}_{0},\hat{y}_{0}) values that are the leading-order terms in Eq. (42) of Theorem 3. Since x^2+y2=O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/3}), root z0z_{0} in the above equation should be nonzero.

We can further show that, near this (x,y)=(12t+x^0,y^0)(x,y)=\left(12t+\hat{x}_{0},\hat{y}_{0}\right) location lies a fundamental lump. This calculation is similar to that we did in the proof of Theorem 2 and the earlier part of this proof of Theorem 3. Specifically, we can show that in the O(1)O(1) neighborhood of this location,

σ(x,y,t)=γ0|T2|NW|3T3/4|d^(d^+1)23|Qd^(z0)|2(|(x12tx^0)+2i(yy^0)+Δ^|2+14)[1+O(|t|1/3)],\displaystyle\sigma(x,y,t)=\gamma_{0}\hskip 1.42271pt|T_{2}|^{N_{W}}|3T_{3}/4|^{\frac{\hat{d}(\hat{d}+1)-2}{3}}\left|Q_{\hat{d}}^{\prime}(z_{0})\right|^{2}\left(\left|(x-12t-\hat{x}_{0})+2\textrm{i}(y-\hat{y}_{0})+\hat{\Delta}\right|^{2}+\frac{1}{4}\right)\left[1+O\left(|t|^{-1/3}\right)\right], (97)

where Δ^=Δ^(Λ,z0)\hat{\Delta}=\hat{\Delta}(\Lambda,z_{0}) is an O(1)O(1) constant. This Δ^\hat{\Delta} is the analog of a similar quantity Δ\Delta which we derived in Eq. (82) for a fundamental lump in the outer region. It is easy to see that the above σ(x,y,t)\sigma(x,y,t) gives a fundamental lump, whose position is at (x,y)=(12t+x0,y0)(x,y)=\left(12t+x_{0},y_{0}\right), where

x0=(z0)(12t)1/3(Δ^),y0=(z0)2(12t)1/312(Δ^),x_{0}=\Re(z_{0})\hskip 1.42271pt(12t)^{1/3}-\Re(\hat{\Delta}),\quad y_{0}=\frac{\Im(z_{0})}{2}(12t)^{1/3}-\frac{1}{2}\Im(\hat{\Delta}), (98)

which matches (42) in Theorem 3. In addition, the error of this prediction is O(|t|1/3)O\left(|t|^{-1/3}\right).

In the center region where x^2+y2=O(1)\hat{x}^{2}+y^{2}=O(1), we can use the technique of Appendix C in Ref. YangYang21a to show that at large time, if zero is a root of the Yablonskii–Vorob’ev polynomial Qd^(z)Q_{\hat{d}}(z), i.e., if d^1mod3\hat{d}\equiv 1\hskip 2.84544pt\mbox{mod}\hskip 2.84544pt3, then uΛ(x,y,t)u_{\Lambda}(x,y,t) would approach a fundamental lump located in the O(1)O(1) neighborhood of the wave center (x^,y)=(0,0)(\hat{x},y)=(0,0). If zero is not a root of Qd^(z)Q_{\hat{d}}(z), then uΛ(x,y,t)u_{\Lambda}(x,y,t) would approach zero in this center region as |t||t|\to\infty. Details will be omitted for brevity. It may be more illuminating for us to point out that, the leading-order term of the previous asymptotic formula (97), which was derived for the region of x^2+y2=O(|t|1/3)\sqrt{\hat{x}^{2}+y^{2}}=O(|t|^{1/3}) and nonzero roots z0z_{0} of Qd^(z)Q_{\hat{d}}(z), turns out to be valid for the x^2+y2=O(1)\hat{x}^{2}+y^{2}=O(1) region and the zero root z0z_{0} of Qd^(z)Q_{\hat{d}}(z) as well [except for the relative error term, which is now O(|t|1)O(|t|^{-1}) rather than O(|t|1/3)O(|t|^{-1/3})]. In other words, if zero is a root of Qd^(z)Q_{\hat{d}}(z), then setting z0=0z_{0}=0 in the leading term of (97), we would get the correct asymptotic fundamental lump in the x^2+y2=O(1)\hat{x}^{2}+y^{2}=O(1) region. In particular, the location of this fundamental lump would be at (x^,y)=(x0,y0)(\hat{x},y)=(x_{0},y_{0}), i.e., (x,y)=(12t+x0,y0)(x,y)=(12t+x_{0},y_{0}), where x0x_{0} and y0y_{0} are given by (98) with z0=0z_{0}=0 and Δ^=Δ^|z0=0\hat{\Delta}=\hat{\Delta}|_{z_{0}=0}. This completes the proof of Theorem 3.

VI Summary and Discussion

In this article, we have analytically studied pattern formation in higher-order lumps of the KP-I equation at large time. For a broad class of these higher-order lumps, we have shown that two types of solution patterns appear at large time. The first type of patterns comprise fundamental lumps arranged in triangular shapes, which are described analytically by root structures of Yablonskii–Vorob’ev polynomials. As time evolves from large negative to large positive, this triangular pattern reverses its xx-direction. The second type of solution patterns comprise fundamental lumps arranged in non-triangular shapes in the outer region, which are described analytically by nonzero-root structures of Wronskian–Hermit polynomials, together with possible fundamental lumps arranged in triangular shapes in the inner region, which are described analytically by root structures of Yablonskii–Vorob’ev polynomials. When time evolves from large negative to large positive, the non-triangular pattern in the outer region switches its xx and yy directions, while the triangular pattern in the inner region reverses its xx-direction. We have also compared these predicted patterns with true solutions, and excellent agreement is observed.

In this pattern analysis of higher-order lumps, we have set the spectral parameter p=1p=1 without any loss of generality (see Remark 4 of Sec. 2). Because of this, lump patterns we have predicted at large time are all yy-symmetric (see Figs. 3, 5 and 7), since root structures of Yablonskii–Vorob’ev and Wronskian-Hermit polynomials are symmetric with respect to the real-zz axis. However, under the Galilean transformation (4), these yy-symmetric lump patterns can become skewed and yy-asymmetric, and these yy-asymmetric patterns correspond to complex spectral parameters pp. Thus, yy-asymmetric lump patterns also exist in the KP-I equation, and such patterns can be obtained from the yy-symmetric ones through the Galilean transformation.

Are there other patterns of higher-order lumps at large time? The answer is yes. Notice that in this article, we have assumed internal-parameter vectors ai\textbf{\emph{a}}_{i} of higher-order lumps to be equal to each other [see Eq. (35)]. If these parameter vectors are allowed to differ from each other, then the analytical results at large time will become different. This problem will not be pursued in this paper, and will be left for future studies.

In a very recent preprint Ling2021 , the authors also derived higher-order lumps in the KP-I equation and studied their large-time patterns through Darboux transformation, and showed that their large-time patterns are described analytically by root structures of Yablonskii–Vorob’ev polynomials. Obviously, the higher-order lump solutions they derived are a very special class of solutions which correspond to the index vector Λ=(1,3,5,,2N1)\Lambda=(1,3,5,\dots,2N-1) and under ai\textbf{\emph{a}}_{i} parameter constraints (35) in our general solutions of Theorem 1, and their large-time pattern results are largely equivalent to our Theorem 2. However, their error estimate of O(|t|2/3)O(|t|^{-2/3}) for fundamental-lump predictions far away from the wave center is different from our O(|t|1/3)O(|t|^{-1/3}) in Theorem 2, and we have verified numerically that their error estimate is incorrect. More importantly, those authors have not considered the more general higher-order lumps corresponding to the index vector Λ(1,3,5,,2N1)\Lambda\neq(1,3,5,\dots,2N-1) in our Theorem 1, nor their large-time solution patterns. These latter patterns are the contents of our Theorem 3 (see also our Figs. 5-8).

Acknowledgment

This material is based upon work supported by the National Science Foundation under award number DMS-1910282, and the Air Force Office of Scientific Research under award number FA9550-18-1-0098.

Appendix

In this appendix, we briefly derive the bilinear higher-order lump solutions presented in Theorem 1.

Under the variable transformation u=2(logτ)xxu=2(\log\tau)_{xx} and notations of x1=x,x2=iyx_{1}=x,x_{2}=\textrm{i}y and x3=4tx_{3}=-4t, the KP-I equation (3) is converted to the bilinear equation

(Dx144Dx1Dx3+3Dx22)ττ=0,(D_{x_{1}}^{4}-4D_{x_{1}}D_{x_{3}}+3D_{x_{2}}^{2})\,\tau\cdot\tau=0, (99)

where DD is Hirota bilinear differential operator. It is well known that if mijm_{ij}, ϕi\phi_{i} and ψj\psi_{j} are functions of (x1,x2,x3)(x_{1},x_{2},x_{3}) and satisfy the following differential equations

x1mij=ϕiψj,\displaystyle\partial_{x_{1}}m_{ij}=\phi_{i}\psi_{j}, (100)
xnϕi=x1nϕi,n=2,3,\displaystyle\partial_{x_{n}}\phi_{i}=\partial_{x_{1}}^{n}\phi_{i},\hskip 42.67912ptn=2,3, (101)
xnψj=(1)n1x1nψj,n=2,3,\displaystyle\partial_{x_{n}}\psi_{j}=(-1)^{n-1}\partial_{x_{1}}^{n}\psi_{j},\quad n=2,3, (102)

then the τ\tau function

τ=det1i,jN(mij)\tau=\det_{1\leq i,j\leq N}\left(m_{ij}\right) (103)

would satisfy the above bilinear equation Hirota_book . To derive higher-order lump solutions, we define mijm_{ij}, ϕi\phi_{i} and ψj\psi_{j} as

mij=𝒜ij1p+qeξi+ηj,ϕi=𝒜ieξi,ψj=jeηj,m_{ij}=\mathcal{A}_{i}\mathcal{B}_{j}\frac{1}{p+q}e^{\xi_{i}+\eta_{j}},\quad\phi_{i}=\mathcal{A}_{i}e^{\xi_{i}},\quad\psi_{j}=\mathcal{B}_{j}e^{\eta_{j}}, (104)

where

𝒜i=1ni!(pp)ni,j=1nj!(qq)nj,\displaystyle\mathcal{A}_{i}=\frac{1}{n_{i}!}(p\partial_{p})^{n_{i}},\quad\mathcal{B}_{j}=\frac{1}{n_{j}!}(q\partial_{q})^{n_{j}}, (105)
ξi=px1+p2x2+p3x3+ξi,0(p),ηj=qx1q2x2+q3x3+ηj,0(q),\xi_{i}=px_{1}+p^{2}x_{2}+p^{3}x_{3}+\xi_{i,0}(p),\quad\eta_{j}=qx_{1}-q^{2}x_{2}+q^{3}x_{3}+\eta_{j,0}(q), (106)

(n1,n2,,nN)(n_{1},n_{2},\cdots,n_{N}) is a vector of arbitrary positive integers, p,qp,q are arbitrary complex constants, and ξi,0(p)\xi_{i,0}(p), ηj,0(q)\eta_{j,0}(q) are arbitrary complex functions of pp and qq. It is easy to see that these mijm_{ij}, ϕi\phi_{i} and ψj\psi_{j} functions satisfy the differential equations (100)-(102). Thus, the above τ\tau function would satisfy the bilinear equation (99). To guarantee that this τ\tau function is real-valued, we impose the parameter constraints

q=p,ηj,0(q)=[ξj,0(p)].q=p^{*},\quad\eta_{j,0}(q)=[\xi_{j,0}(p)]^{*}. (107)

Under these constraints, ηj=ξj\eta_{j}=\xi_{j}^{*}, mni,nj=mnj,nim_{n_{i},n_{j}}^{*}=m_{n_{j},n_{i}}, and thus τ\tau in (103) is real. Following the technique of Ref. OhtaJY2012 , we can further show this τ\tau is positive. Thus, the resulting function u=2(logτ)xxu=2(\log\tau)_{xx} is a real-valued solution to the KP-I equation (3).

Next, we need to simplify the matrix elements of this τ\tau determinant and derive their more explicit algebraic expressions. This simplification is very similar to that we performed in OhtaJY2012 ; YangYang3wave . By expanding ξi,0(p)\xi_{i,0}(p) into a certain series containing complex parameters ai=(ai,1,ai,2,)\textbf{\emph{a}}_{i}=\left(a_{i,1},a_{i,2},\cdots\right) and repeating the calculations of OhtaJY2012 ; YangYang3wave , we can show that the matrix element mijm_{ij} in (104) can be reduced to the expression given in Eq. (9) of Theorem 1.

We would like to make a comment here regarding the choice of differential operators in Eq. (105). Obviously, we can also choose more general forms of these differential operators, such as

𝒜i=1ni![f(p)p]ni,j=1nj![f(q)q]nj,\displaystyle\mathcal{A}_{i}=\frac{1}{n_{i}!}\left[f(p)\partial_{p}\right]^{n_{i}},\quad\mathcal{B}_{j}=\frac{1}{n_{j}!}\left[f(q)\partial_{q}\right]^{n_{j}}, (108)

where f(p)f(p) is an arbitrary function, and the resulting τ\tau function (103) would still satisfy the bilinear equation (99). However, such additional freedoms in the differential operators will not produce new higher-order lump solutions. To see why, we can rewrite this 𝒜i\mathcal{A}_{i} as

𝒜i=1ni![f(p)ppp]ni=k=0nici,k1(nik)!(pp)nik,\mathcal{A}_{i}=\frac{1}{n_{i}!}\left[\frac{f(p)}{p}p\partial_{p}\right]^{n_{i}}=\sum_{k=0}^{n_{i}}c_{i,k}\frac{1}{(n_{i}-k)!}(p\partial_{p})^{n_{i}-k}, (109)

where ci,kc_{i,k} are pp-dependent complex constants. Similar treatments can be made on j\mathcal{B}_{j}. These differential operators in summation form are similar to those taken in Ref. OhtaJY2012 . We can directly show that the mijm_{ij} matrix element with these differential operators of summation form can be converted to one with these differential operators as a single term in (105), after parameters ai\textbf{\emph{a}}_{i} in the series expansion of ξj,0(p)\xi_{j,0}(p) are redefined properly. Thus, no new solutions are produced.

References

References

  • (1) B.B. Kadomtsev and V.I. Petviashvili, “On the stability of solitary waves in weakly dispersive media”, Sov. Phys. Dokl. 15, 539-541 (1970).
  • (2) M.J. Ablowitz and H. Segur, “On the evolution of packets of water waves”, J. Fluid Mech. 92, 691-715 (1979).
  • (3) D.E. Pelinovsky, Yu.A. Stepanyants and Yu.A. Kivshar, “Self-focusing of plane dark solitons in nonlinear defocusing media”, Phys. Rev. E 51, 5016-5026 (1995).
  • (4) S. Tsuchiya, F. Dalfovo and L.P. Pitaevskii, “Solitons in two-dimensional Bose-Einstein condensates”, Phys. Rev. A 77, 045601 (2008).
  • (5) S. Novikov, S.V. Manakov, L.P. Pitaevskii and V.E. Zakharov, Theory of solitons: the inverse scattering method (Plenum, 1984).
  • (6) M.J. Ablowitz and P.A. Clarkson, Solitons, Nonlinear Evolution Equations and Inverse Scattering (Cambridge University Press, 1991).
  • (7) V.I. Petviashvili, “Equation of an extraordinary soliton”, Plasma Physics 2, 469-472 (1976).
  • (8) S.V. Manakov, V.E. Zakharov, L.A. Bordag, A.R. Its and V.B. Matveev, “Two-dimensional solitons of the Kadomtsev–Petviashvili equation and their interaction”. Phys. Lett. A 63, 205-206 (1977).
  • (9) J. Satsuma and M.J. Ablowitz, “Two-dimensional lumps in nonlinear dispersive systems”, J. Math. Phys. 20, 1496 (1979).
  • (10) K. A. Gorshkov, D. E. Pelinovsky, and Yu. A. Stepanyants, “Normal and anomalous scattering, formation and decay of bound states of two-dimensional solitons described by the Kadomtsev–Petviashvili equation,” JETP 77, 237-245 (1993).
  • (11) M.J. Ablowitz and J. Villarroel, “Solutions to the time dependent Schrödinger and the Kadomtsev-Petviashvili equations,” Phys. Rev. Lett. 78, 570-573 (1997).
  • (12) M.J. Ablowitz, S. Chakravarty, A.D. Trubatch and J. Villarroel, “A novel class of solutions of the non-stationary Schrödinger and the Kadomtsev–Petviashvili I equations”, Phys. Lett. A 267, 132-146 (2000).
  • (13) D.E. Pelinovsky and Yu. A. Stepanyants, “New multisoliton solutions of the Kadomtsev–Petviashvili equation”, JETP Lett. 57, 24-28 (1993).
  • (14) D. Pelinovsky, “Rational solutions of the KP hierarchy and the dynamics of their poles. II. Construction of the degenerate polynomial solutions”, J. Math. Phys. 39, 5377-5395 (1998).
  • (15) P. Dubard, P. Gaillard, C. Klein and V.B. Matveev, “On multi-rogue wave solutions of the NLS equation and positon solutions of the KdV equation”, Eur. Phys. J. Spec. Top. 185, 247-258 (2010).
  • (16) P. Dubard and V.B. Matveev, “Multi-rogue waves solutions: from the NLS to the KP-I equation”, Nonlinearity 26, R93-R125 (2013).
  • (17) S. Chen, P. Grelu, D. Mihalache and F. Baronio, “Families of rational solution solutions of the Kadomtsev–Petviashvili I equation”, Rom. Rep. Phys. 68, 1407-1424 (2016).
  • (18) P.A. Clarkson and E. Dowie, “Rational solutions of the Boussinesq equation and applications to rogue waves”, Trans. Math. Appl. 1, 1-26 (2017).
  • (19) P. Gaillard, “Multiparametric families of solutions of the Kadomtsev–Petviashvili-I equation, the structure of their rational representations, and multi-rogue waves”, Theor. Math. Phys. 196, 1174-1199 (2018).
  • (20) J.H. Chang, “Asymptotic analysis of multilump solutions of the Kadomtsev–Petviashvili-I equation”, Theor. Math. Phys. 195, 676-689 (2018).
  • (21) W.X. Ma, “Lump solutions to the Kadomtsev–Petviashvili equation”, Phys. Lett. A 379, 1975-1978 (2015).
  • (22) C. Lester, A. Gelash, D. Zakharov and V.E. Zakharov, “Lump chains in the KP-I equation”, Stud. Appl. Math. (2021), doi.org/10.1111/sapm.12420 (see also arXiv:2102.07038).
  • (23) J. Rao, K.W. Chow, D. Mihalache and J.S. He, “Completely resonant collision of lumps and line solitons in the Kadomtsev–Petviashvili I equation”, Stud. Appl. Math. 2021 (doi.org/10.1111/sapm.12417).
  • (24) B. Yang and J. Yang, “Rogue wave patterns in the nonlinear Schrodinger equation”, Physica D 419, 132850 (2021).
  • (25) B. Yang and J. Yang, “Universal rogue wave patterns associated with the Yablonskii-Vorob’ev polynomial hierarchy”, Physica D 425, 132958 (2021).
  • (26) J. Weiss, “Modified equations, rational solutions, and the Painlevé property for the Kadomtsev–Petviashvili and Hirota–Satsuma equations”, J. Math. Phys. 26, 2174 (1985).
  • (27) Y. Ohta, J. Yang, “General high-order rogue waves and their dynamics in the nonlinear Schrödinger equation”, Proc. Roy. Soc. A 468, 1716 (2012).
  • (28) B. Yang and J. Yang, “General rogue waves in the three-wave resonant interaction systems”, IMA J. Appl. Math. 86, 378-425 (2021).
  • (29) B. Yang and J. Yang, “General rogue waves in the Boussinesq equation”, J. Phys. Soc. Jpn. 89, 024003 (2020).
  • (30) J. Chen, Y. Chen, B.F. Feng, K. Maruno and Y. Ohta, “General high-order rogue waves of the (1+1)-dimensional Yajima–Oikawa system”, J. Phys. Soc. Jpn. 87, 094007 (2018).
  • (31) A.I. Yablonskii, Vesti Akad. Navuk. BSSR Ser. Fiz. Tkh. Nauk. 3, 30 (1959) (in Russian).
  • (32) A.P. Vorob’ev, “On rational solutions of the second Painlevé equation”, Diff. Eqns. 1, 58 (1965).
  • (33) K. Kajiwara and Y. Ohta, “Determinant structure of the rational solutions for the Painlevé II equation”, J. Math. Phys. 37, 4693 (1996).
  • (34) P.A. Clarkson and E.L. Mansfield, “The second Painlevé equation, its hierarchy and associated special polynomials”, Nonlinearity 16, R1 (2003).
  • (35) S. Fukutani, K. Okamoto and H. Umemura, “Special polynomials and the Hirota bilinear relations of the second and the fourth Painlevé equations”, Nagoya Math. J. 159, 179-200 (2000).
  • (36) M. Taneda, “Remarks on the Yablonskii-Vorob’ev polynomials”, Nagoya Math. J. 159, 87-111 (2000).
  • (37) R.J. Buckingham and P.D. Miller, “Large-degree asymptotics of rational Painlevé-II functions: noncritical behaviour”, Nonlinearity 27, 2489 (2014).
  • (38) F. Balogh, M. Bertola and T. Bothner, “Hankel determinant approach to generalized Vorob’ev-Yablonski polynomials and their roots”, Constr. Approx. 44, 417 (2016).
  • (39) G. Felder, A.D. Hemery and A.P. Veselov, “Zeros of Wronskians of Hermite polynomials and Young diagrams”, Physica D 241, 2131-2137 (2012).
  • (40) M. Garca-Ferrero and D. Gmez-Ullate, “Oscillation theorems for the Wronskian of an arbitrary sequence of eigenfunctions of Schrödinger’s equation”, Lett. Math. Phys. 105, 551-573 (2015).
  • (41) N. Bonneux, C. Dunning and M. Stevens, “Coefficients of Wronskian Hermite polynomials”, Stud. Appl. Math. 144, 245–288 (2020).
  • (42) J. Dong, L. Ling and X. Zhang, “Kadomtsev-Petviashvili equation: one-constraint method and lump pattern”, arXiv:2108.09715 [nlin.SI] (2021).
  • (43) R. Hirota, The direct method in soliton theory (Cambridge University Press, Cambridge, 2004).