Blowup for the defocusing septic complex-valued nonlinear wave equation in
Abstract.
In this paper, we prove blowup for the defocusing septic complex-valued nonlinear wave equation in . This work builds on the earlier results of Shao, Wei, and Zhang [56, 55], reducing the order of the nonlinearity from to in . As in [56, 55], the proof hinges on a connection between solutions to the nonlinear wave equation and the relativistic Euler equations via a front compression blowup mechanism. More specifically, the problem is reduced to constructing smooth, radially symmetric, self-similar imploding profiles for the relativistic Euler equations.
As with implosion for the compressible Euler equations, the relativistic analogue admits a countable family of smooth imploding profiles. The result in [56] represents the construction of the first profile in this family. In this paper, we construct a sequence of solutions corresponding to the higher-order profiles in the family. This allows us to saturate the inequalities necessary to show blowup for the defocusing complex-valued nonlinear wave equation with an integer order of nonlinearity and radial symmetry via this mechanism.
1. Introduction
We investigate finite time blowup for the complex-valued defocusing nonlinear wave equation:
(1.1) |
for . Given smooth and well-localized initial data , the classical Cauchy theory states that (1.1) admits a smooth and strong local-in-time solution [61]. The parameters , can be naturally classified in terms of the critically of the conserved energy
in terms of the scaling symmetry
Precisely, we obtain three regimes: the subcritical case, or for ; the critical case, and ; and the supercritical case, and .
The blowup of nonlinear wave equations has been extensively studied in the focusing case (cf. [3, 46, 38, 45, 24, 25, 26, 27, 37, 40, 41, 42, 43]). For the defocusing case, the equation is globally well-posed in the subcritical and critical case [28, 29, 32, 31, 57, 58, 39, 63]. Nevertheless, the question of blowup or global well-posedness for smooth solutions to defocusing nonlinear Schrödinger or wave equation has remained a long-standing open problem [65]. Tao, in [64], showed if one considers a supercritical, high dimensional, defocusing nonlinear wave system , for some smooth, positive, carefully designed potential , one proves blowup in finite time. Recently, in the breakthrough work of Merle, Raphael, Rodnianski and Szeftel [50], blowup for the supercritical defocusing nonlinear Schrödinger equation was proven via a novel front compression mechanism, whereby the phase of the solution blows up. The work [49] leverages a connection between the nonlinear Schrödinger equations and the compressible Euler equations with an added quantum pressure term via Madelung’s transformation. Under the appropriate scaling assumptions – which necessitates considering dimensions and high order nonlinearities – the quantum pressure term becomes lower order, which enabled the use of self-similar imploding profiles to the compressible Euler equations constructed in [51], as asymptotic self-similar profiles to the defocusing nonlinear Schrödinger equation.
Inspired by the work [50], Shao, Wei, and Zhang in [56, 55] recently exploited a similar front compression mechanism to establish blowup for the defocusing nonlinear wave equation (1.1) via a self-similar profile to the relativistic Euler equations. Specifically, they proved finite time blowup of (1.1) from smooth compactly supported initial data with dimension and odd nonlinearity satisfying and . See also Theorem 1.2.
1.1. Self-similar ansatz
We consider the phase-amplitude representation with being real and rewrite (1.1) as the system of . To this end, we compute
Substituting the above computation into (1.1), dividing by and then taking the real and imaginary part of the equation, we derive the equations for and
(1.2a) | ||||
(1.2b) |
We consider the general self-similar blowup ansatz
(1.3) |
for some to be determined.
In line with front compression, where the phase blows up, we consider for our self-similar ansatz (1.3). Moreover, to obtain a collapsing blowup, which is natural due to the conservation of the energy, we impose . We use the notation to denote that and are comparable up to a independent constant. Under these assumptions, for , we get
(1.4) |
Since , we treat in (1.2a) as lower order terms. For the -equation (1.2b), it is easy to observe that the scalings of each terms are balanced. Thus, dropping in (1.2a) and then dividing on both sides of (1.2a), we derive the following leading order system of (1.2a)
(1.5) |
To balance these three terms, using (1.4), we impose the following relation among in (1.3)
(1.6) |
Note that equations (1.5) and (1.2b) reduce to isentropic relativistic Euler equation. In fact, by introducing , the flat Lorentz metric and derivative as
the relativistic velocity , the energy density , the energy momentum tensor , and the pressure subject to the isothermal equation of state as follows
(1.7a) | |||
we derive the relativistic Euler equations | |||
(1.7b) |
with normalization . The above connection between the defocusing wave equations and the relativistic Euler equations has been used in [56].
The local existence of smooth solutions to the relativistic Euler equations was established in [48], with some global existence results in [60, 54, 14]. Meanwhile, the equations can develop a finite time singularity [35, 53, 21]. For more discussions on the mathematical aspects of relativistic fluids, see the surveys [23, 2].
1.2. Main results
By imposing the radial symmetry on and plugging the self-similar ansatz (1.3) in equations (1.2b) and (1.5), we can derive the ODE (2.3) for the self-similar profiles (1.3). See the derivation in Section 2.1.
To use the front compression mechanism for smooth blowup, we require to satisfy
(1.8) |
which implies and the non-linear exponent . For and integer , this implies . See Section 2.1 for more details about these constraints. In this paper, we consider the lowest dimension in this setting and the smallest integer exponent in this dimension.
Our main result is stated as follows.
Theorem 1.1.
Let , and be an odd number and large enough. There exists accumulating at with such that the ODE (2.3) admits a smooth solution with ,
(1.9) |
for any , and for some function . Moreover, the solution gives rise to a smooth radially symmetric self-similar profile in (1.3) and a solution to the relativistic Euler equations (1.2b) and (1.5) on :
(1.10) |
where . The solution develops an imploding singularity at and satisfies the asymptotics
(1.11) |
for any and some constants
Theorem 1.2 (Theorem 1.1 [55]).
Let with . Suppose that there exists with and 111 The parameter in [55] corresponds to (1.3), (2.1) and in this paper. The condition in [55, Theorem 1] is equivalent to and (2.4). such that the ODE (2.3) with parameters admits a smooth solution defined in satisfying and for some function .
Then there exists smooth compactly supported functions such that the defocusing nonlinear wave equation (1.1) with nonlinearity develops a finite time singularity from the initial data .
In Theorem 1.1, since and converges to as , for large enough, the parameters satisfies the inequalities in Theorem 1.2. Moreover, the smooth solution constructed in Theorem 1.1 satisfies the assumptions in Theorem 1.2. Thus, using Theorems 1.1 and 1.2, we establish:
Corollary 1.3.
There exists smooth compactly supported functions such that the septic defocusing nonlinear wave equation (1.1) with develops a finite time singularity from the initial data .
Remark 1.4.
In dimension with nonlinearity , the range of parameter in the supercritical case implies that and . Our result resolves the case . It is conceivable that our methods for proving Theorem 1.1 can apply to a specific pair of parameters with , provided they satisfy (1.8). Note that the parameters in the supercritical case with and that do not satisfy (1.8) are .
The main difficulty to prove Theorem 1.1 is that the ODE (2.3) of is singular near a sonic point. Additionally, the ODE degenerates near the sonic point as . We remark that, due to the constraint (1.8), constructing the blowup solution to (1.1) with smaller requires considering close to . To overcome these difficulties, we renormalize the ODE (2.3) to obtain a new ODE of (2.14), which is non-degenerate near the sonic point as . To analyze the -ODE near the sonic point, we perform power series expansion of and estimate the asymptotics of the power series coefficients using induction. Using barrier arguments, a shooting argument, and a few monotone properties, we extend the local power series solution to a global solution of the ODE.
The proof involves light computer assistance, mainly to derive the power series coefficients for and to verify the signs of a few lower-order polynomials. These calculations can be performed on a personal laptop in a few seconds. See more details in Appendix B.
1.3. Comparison with the methods in existing works
Our proof share some similarities with [6], including some barrier arguments, estimates of the power series coefficients using induction, and computer-assisted proofs. Different from [6], we need to renormalize the ODE to overcome the degeneracy as .
In [6], the estimates of the asymptotics of power series coefficients involve ad-hoc partitions of . We develop a systematic approach for the estimates by identifying a few leading order terms in the recursive formula of and treating the remaining terms perturbatively. We check a few properties of for suitably large with computer assistance and then perform induction for the remaining coefficients. This allows us to impose a stronger induction hypothesis, resulting in a streamlined approach to estimating the higher order coefficients, which further leads to a simplified splitting in . Moreover, our scheme of estimates does not utilize special forms of the ODE and are readily generalized to ODE (2.3) or (2.14) with numerator and denominator being higher order polynomials. Additional simplifications of the combinatorial estimates are made via renormalization of the power series coefficients. See Section 3.
We follow the connection between the defocusing nonlinear wave equation and the relativistic Euler equation used in [56] to derive the ODE (2.3) governing the profiles, use the same renormalization as in [56], and adopt a few basic derivations for the ODEs from [56]. However, there are a few key differences. Firstly, we apply renormalization to overcome the degeneracy of the ODE as , whereas in [56], renormalization is used to simplify computations and obtain blowup results for a wider range of . We can also employ other renormalization to overcome this degeneracy; see Remark 2.3. We adopt the renormalization from [56] to simplify certain presentations and derivations that are less essential. Secondly, since we consider a smaller nonlinearity () than those considered in [56], we need to use much higher-order barrier functions near the sonic point than those in [56], which are based on quartic or lower-order polynomials. Thirdly, by taking , we obtain a large parameter (2.30) and use it and light computer assistance to bypass several detailed computations in [56].
We note that in [34], Guo, Hadzic, Jang, and Schrecker employed similar arguments, such as Taylor expansions, dynamical systems analysis, and computer-assisted proofs, to construct smooth self-similar solutions for the gravitational Euler-Poisson system.
1.4. Related results
In this section, we review related results on singularity formation and computer-assisted proofs in fluid mechanics
Smooth Implosions
The construction of the self-similar imploding singularity in Theorem 1.1 is closely related to that in compressible fluids. Guderley [33] was the first to construct radial, non-smooth, imploding self-similar singularities along with converging shock waves in compressible fluids. While Guderley’s setting has been extensively studied, the existence of a finite-time smooth, radially symmetric implosion (without shock waves) was only recently established in [51, 52, 50]. Subsequently, radially symmetric implosion in with a larger range of adiabatic exponents was established in [6], and non-radial implosion was established in [12, 11]. While these results are inspired by Guderley’s setting, constructing self-similar implosion profiles is challenging due to the degeneracy at the sonic point, requiring sophisticated mathematical techniques [51] or computer-assisted methods [6]. The smooth imploding solutions are proved to be potentially highly unstable [52, 6, 12, 50], with numerical studies of instabilities presented in [5]. The self-similar profile in Guderley’s setting was recently constructed in [36]. By perturbing the radial implosion [51], exploiting axisymmetry, and proving the full stability of the perturbation generated by angular velocity, vorticity blowup in the compressible Euler equations was established in [16] for the case of and in [15] for with .
Shock formation
The prototypical singularity for the compressible Euler equations is the development of a shock wave. Rigorous work regarding shocks traces back to the work of Lax [44]. Shock wave singularities in the multi-dimensional, irrotational, isentropic setting was first analyzed by Christodoulou in the work [21] (cf. [22]). This work was later extended to include non-trivial vorticity in the 2-dimensional setting by Luk and Speck [47]. Restricting to 2-dimensions and assuming azimuthial-symmetry, the first complete description of the formation of a shock singularity, including its self-similar structure was given in a work of the first author, Shkoller, and Vicol [8]. This latter work was extended by the authors to 3-dimensions in the absence of symmetry assumptions in [9] and to the 3-dimensional non-isentropic setting in [10]. Returning to the setting of 2-dimensions under azimuthial-symmetry, the first full description of shock development past the first singularity was proven in a work by the first author, Drivas, Shkoller, and Vicol [7]. Shock development in the absence of symmetry remains an open problem; however, recently Shkoller, and Vicol in the work [59] constructed compact in time maximal development in the 2-dimensional setting, which is a crucial step towards resolving this major open problem (partial results in direction of maximal development were earlier attained in [1]).
Computer-assisted proofs in fluids
In recent years, there have been substantial developments in computer-assisted proofs in mathematical fluid mechanics. We highlight a few advancements in incompressible fluids: singularity formation in the 3D Euler equations [18, 17] and related models [19, 20], constructing nontrivial global smooth solutions to SQG [13], and some applications to Navier-Stokes [4, 66]. See also the survey [30].
2. Autonomous ODE for the self-similar profile
In this section, we derive the ODE governing the self-similar profile (1.3) and perform renormalization to overcome some degeneracy.
2.1. ODE for the profile
We introduce and a free parameter to rewrite in (1.3), (1.6) as follows
(2.1) |
where is the dimension. We impose radial symmetry on the solution . As a result, the profiles in (1.6) are radially symmetric. We introduce the radial and self-similar variables
(2.2a) | |||
From the ansatz (1.3) and (1.6), it is easy to see that only depend on . Thus, to solve (1.5), we consider | |||
(2.2b) |
The relation (2.2) and (1.2b) imply that solves the following ODE for
(2.3) | ||||
This ODE was first derived in [56]. For completeness, we derive the ODE in Appendix A.1. 222 We have substitute in the ODEs derived in [56, Section 2] to reduce the number of variables. Moreover, we have used to remove the parameter used in [56]. Note that the meaning of is different in [56] and [55].
Constraints on the parameters
From (1.6), which is related to the front compression mechanism and the relation (2.1), we first require to satisfy
(2.4) |
To obtain a smooth ODE solution, we have an additional constraint on established in [56, Lemma 2.1].
Lemma 2.1 (Lemma 2.1 [56]).
Since , we need . Recall the power of non-linearity (2.1). Let be the value with . For , the constraint (2.5) implies
In the remainder of the paper, we will focus on the lowest dimension and the smallest integer power in this dimension. Then, we have
(2.6) |
We still have one free parameter , which will be chosen so that .
Double roots and the sonic point
Solving , we get the special points
(2.7) |
where is the sonic point. The letters are short for origin, sonic, respectively.
The smoothness of the ODE solution to (2.3) is closely related to the Jacobian of a renormalization of the ODE (2.3) by the factor :
(2.8) |
A direct computation shows that the entries degenerate: , as , which complicates the analysis of the ODE near the sonic point.
Notations and parameters
Throughout the paper, we will use for the variables in the original ODE system (2.3), and for the renormalized ODE (2.14) to be introduced. We further introduce some parameters
(2.9) |
and will use them to denote the coefficients of a polynomial in the renormalized ODE of .
We fix the following constants
(2.10) |
and will use them to denote the size of relative error for asymptotic series around the sonic point (see Lemma 3.5 and Corollary 3.11).
We denote if for some absolute constant , and denote if and . We denote if for some constant depending on , and define the notations similarly.
2.2. Renormalization
To overcome the degeneracy of the Jacobian (2.8) as , we study an ODE system equivalent to (2.3) by performing a renormalization of the system. We adopt the change of coordinate from [56] with
(2.11) |
We will see later that the Jacobian of the new ODE system in at the sonic point (2.26) is non-degenerate as . There are a few other change of coordinates that achieve this purpose. See Remark 2.3. The above transform (2.11) leads to lower order polynomials in , which simplifies some analysis. Note that we do not use this property in an essential way.
We can invert the transform from to using the following formulas
(2.12) |
The bijective properties are proved in [56, Lemma 3.8] :
Lemma 2.2.
Denote . Then is a bijection.
A direct computation yields that the inversion of is given by . Thus, for or , we have
(2.13) |
Suppose that solves (2.3). We have the following ODE for :
(2.14) | ||||
where are defined in (2.9). We refer to [56, Section 3.2] for the derivations of (2.14).
Denote by the degree of as a polynomial in , respectively. We have
(2.15a) | |||
We can expand as polynomials in | |||
(2.15b) | |||
where are polynomials in . We further define the maximum degree of these polynomials in as follows | |||
(2.15c) |
For the above ODE system (2.14), we have
(2.15d) |
In this new coordinate system , the points defined in (2.7) map to
(2.16) |
Remark 2.3.
One can choose other change of variables to make the matrix (2.8) nonsingular and analyze the ODE in the new system. One candidate is the following simpler transform
In the new system, the gradient at the sonic point does not degenerate in the limit
The reason is that the transformation from to is singular near as with , which compensates the degeneracy: .
2.2.1. Change of coordinates
Recall the maps from to (2.11) and from to (2.12). Denote by the Jacobians
(2.17) |
For with defined in Lemma 2.2, the denominators in the maps (2.11), (2.12) do not vanish. Thus, the matrices are not singular. Using (2.13), we have
(2.18) |
Since is not singular, we get .
Along the ODE solution curve or , the ODE (2.14) and the above change of coordinates implies
where denotes the -th component of a matrix . Thus, there exists some continuous function such that
(2.19) |
for or equivalently . For , since and , we obtain .
2.3. Roots of
We can decompose , defined in (2.3) as follows
(2.20) | ||||
The derivations follow from a direct computation. See Figure 1 for an illustration of the curves (red) and (blue). For , a direct computation yields
(2.21) |

Plugging the formulas of (2.12) in (2.3) with , we can also rewrite as
(2.22) |
We refer to [56, Section 3.2] for the derivation of (2.22). The numerator can be written as , where we define
(2.23) |
We will use the function in Section 5 to control the sign of on the solution curve.
We can decompose (2.14) as follows
(2.24a) | |||
with given by | |||
(2.24b) |
See Figure 2 for an illustration of the curves (red), (blue).
We have the following basic properties about .
Lemma 2.4.
Suppose that satisfies . For , we have . For , we have .
The above properties hold for a wider range of parameters , but we restrict the range to simplify the proof. The parameters in (2.6) with satisfy the above assumptions.
2.4. Eigen-system near
We define
(2.26a) | ||||||
Using the definitions in (2.14) and a direct computation, at , we yield 444 These formulas have also been derived in [56, Sections 4.1, 4.2]. | ||||||
(2.26b) | ||||||
where are defined in (2.10). The eigenvalues of are given by
(2.27a) | ||||
Using the above formulas of and (2.6), (2.9), we compute | ||||
(2.27b) |
We view all the parameters as functions in . From (2.27), for , we obtain
(2.28) |
Since (2.9), using the above estimates and (2.9), we obtain
(2.29) |
We introduce the parameter
(2.30) |
related to the asymptotics of the power series of near the sonic point. Using (2.27), we obtain
(2.31) |
We want to study the smooth self-similar profile with sufficiently large, which provides an important large parameter in our analysis. Since , the numerator is always larger than . Thus, we want to choose with close to
(2.32) |
which is consistent with the constraint in Lemma 2.1. Note that given , we can determine via (2.31). It is not difficult to see that decreases in for with some absolute constant , is a smooth bijection, and it admits an inverse map :
(2.33) |
Thus, for each integer , there exists strictly decreasing such that
(2.34) |
In the remainder of the paper, we drop the dependence of on for simplicity. Whenever we refer to choosing sufficiently large, it is equivalent to taking sufficiently close to .
Next, we use the ODE (2.15b) to compute the slope of the solution at the sonic point , which is given by with
Denote . Applying Taylor expansion near (2.26), we get
(2.35) |
Using the derivation for from (2.26) and the ODE (2.15b), we yield
(2.36) |
It is not difficult to show that (2.36) implies that is an eigenfunction of (2.26). Moreover, using (2.26b) and (2.27b), we get . Thus, the above equation has two roots . The eigenfunction associated with eigenvalue is parallel to
We want to construct a smooth curve that pasts through along the positive direction . See the black curve in Figure 2 for an illustration of the direction. Using (2.27), (2.26), (2.6), (2.9), for close to , which implies , we get that the direction corresponds to . We fix . Since and are parallel, we can represent and (2.31) as follows
(2.37) |
2.5. Outline of the proof
In this section, we outline the proof of Theorem 1.1.
In Section 3, we analyze the ODE (2.14) near the sonic point (2.16) by constructing power series solution and estimating the asymptotics of the power series coefficients using induction.
In Section 4, we use a double barrier argument, a shooting argument, to extend the local power series solution to for some small , which along with a gluing argument gives rise to a smooth ODE solution to (2.3) for with small . It corresponds to a solution curve connecting . See the black curve in Figure 1 or the curves above in Figure 2 for illustrations.
3. Power series near the sonic point
In this section, we study the behavior of the ODE (2.14) near the sonic point (2.16) using power series expansion
(3.1) |
Although we have in our case, we aim to develop a general method which does not depend on fine properties of the ODEs, e.g., specific value or degree of the polynomials.
3.1. Recursive formula
In this section, our goal is to establish Lemma 3.2 for the recursive formula of the power series coefficients of the ODE solution near the sonic point .
Notations
Throughout this section, we use to denote variables, and to denote the value of the power series coefficients of the solution to the ODE (2.14).
Firstly, for any two power series with coefficients , i.e. , we get
(3.2) |
Given any power series with coefficients and (3.1), since are polynomials of , using the above convolution formula, we can obtain the power series of
(3.3) |
Applying chain rule, one can obtain that only depends on . For example, we have the following formula for the first term
(3.4) |
We introduce the following functions
(3.5) |
It is easy to see that only depends on using the chain rule. We will use them in Lemmas 3.1, 3.2. For simplicity, we drop the dependence of on .
We have the following formula regarding the coefficients of .
Lemma 3.1.
Proof.
3.1.1. Derivation of the recursive formulas
Now, we derive the recursive formula for the power series coefficients of (3.1) near :
(3.8) |
with satisfying the ODE (2.3).
Firstly, recall the notations from (2.16) and (2.26). We have and
(3.9) |
Thus, the leading coefficient satisfies for . Using the ODE (2.3) together with the expansions (3.3) and (3.8), we obtain
Matching the coefficients of , we get
(3.10) |
where evaluate on .
Let us define
(3.11) |
From (3.10), evaluating at (), we have
(3.12) |
Next, we derive another formula of by determining the explicit dependence on the terms . Then, by evaluating on , we can obtain the recursive formula of in terms of for . We will choose
(3.13) |
When we use the recursive formula later, e.g., in Section 3.2, we choose with given in (3.23).
Expansion of
For with , taking derivative on (3.11) and using (3.6c), we obtain
and then using (3.6a) and changing , we get
These identities hold for functions evaluated in any .
For with and (see the bound in (3.13)), we have . Thus, the right-hand side of the above formula depends only on and is independent of , and depends linearly on .
To simplify the notations, we introduce the following functions
(3.14a) | ||||
where are defined in (3.5). Then we can rewrite as follows | ||||
(3.14b) |
The above discussion implies the following expansion of in the top terms
(3.15) |
Lemma 3.2.
Given , for , the coefficients of the power series (3.1) solving (2.3) near satisfy
(3.16) | ||||
where we abuse notation by denoting the value of the functions (3.14) evaluating on for . For with , only depend on . Here, we use the convention
if . In particular, for and , we have
(3.17) |
where are defined in (2.27a) and the right hand side is independent of .
Recall from (2.26). With , we can further update as follows
(3.18) | ||||
Remark 3.3 (Parameters and coefficients).
Proof.
Evaluating (3.15) on for any , and applying (3.10) and the definition (3.11) to , we prove the identity (3.16). From (3.3), we know that only depends on .
Note that the right hand side of (3.17) is the same as (3.16) with . Thus to prove (3.17), we only need to evaluate on . We first compute . Using the definition (3.5) and the notations (2.26), we have
(3.19) |
Using (3.14) and the formulas of (3.19), (3.4), and then (2.37), we obtain
and prove (3.17). Using (3.6a) with , we obtain
From (3.19), we obtain that is independent of (3.19). Thus, we prove (3.18).
To simplify the estimate of the power series coefficients, we consider the renormalized coefficient
(3.20a) | |||
where is the Catalan number and satisfies the following asymptotics and identity | |||
(3.20b) | |||
for any . See [62, Section 1] for the identity. Moreover, for , we yield | |||
(3.20c) |
3.2. Asymptotics of the coefficients
In this Section, we develop refine estimates of , which are crucial for the barrier argument Sections 4, 5. Although the relations (3.16) and (3.17) seem to be complicated, we will show that the top two terms and are the dominated terms and determine the asymptotics of . Other terms can be treated perturbatively. Using Lemma 3.1 and (3.14), one can show that
(3.21) |
We expect that the asymptotic growth rate of is given by the above rate and will justify it in (3.22e) in Lemma 3.5 and Corollary 3.11, equation (3.61).
We have the following estimate for binomial coefficients.
Lemma 3.4.
For any and , we have .
The proof is elementary and we refer to [6, Lemma 5.5].
We have the following estimates of the asymptotics of . 555In Lemma 3.5 and its proof, we do not assume that .
Lemma 3.5.
Let be the parameter chosen in (3.23). There exists large enough such that for any and , the following statements hold true.
For any , we get 666 In the upper bound of (3.22a), we use instead of the more symmetric form since can be . In our case, we have (2.16) as . | |||
(3.22a) |
For any , we have
(3.22b) |
For any , we have
(3.22c) |
Let be chosen in (2.10), and denote
(3.22d) |
For any with , we get
(3.22e) |
In (3.22a), (3.22c), (3.22e), we consider the range of : larger than for later estimates in Section 4. From (3.22e), for , changes sign, and (3.22b) does not hold.
Ideas and strategy
We expect that the asymptotics of is given by (3.22e), which quantifies the growth rate (3.21) for . Since grows very fast in , we can treat the remaining terms in (3.17), (3.16) perturbatively.
Recall from (2.15a). To prove the induction, we introduce a few parameters
(3.23) |
We require in (3.22b)-(3.22e) with not too small, since these estimates may not hold true for small . We list a few conditions satisfying by the parameters (3.23) in the following lemma. We verify them with computer assistance and will use them to prove Lemma 3.5. See more discussions on the computer assistance in Appendix B.
Lemma 3.6 (Computer-assisted).
Let be the parameters in (3.24), in (3.22d), and in (3.22a). For (thus (2.31)), the following statements hold true.
(b) For any with , we have
(3.24) |
(c) The parameter defined below satisfies
(3.25) |
Let us motivate the choice of the parameters in (3.23). First, we fix as in (3.23). Next, we choose large enough so that conditions (3.22e), (3.22c), and (3.22b) begin to hold. We then choose large enough to verify (3.25). With fixed, (3.24) holds for any with . We choose large and then take to be sufficiently large. See the discussion below (3.47).
Using (3.17) and (2.29), we can obtain that is continuous in as for any . With Lemma 3.6 and continuity of in , Lemma 3.5 holds for all and large enough (equivalent to small). Below, we assume that the inductive hypothesis is true for the case of and we will prove the case of with . In Section 3.3, we derive the consequence of the inductive hypothesis. In Section 3.4, we estimate the coefficients of the power series of . In Section 3.5.4, we estimate and prove the induction.
Remark 3.7.
The proof presented below does not take advantage of the specific forms of , which are polynomials in with low degree . We can prove it by checking the desired properties of with large enough with computer assistance and then use induction to prove the properties of with large enough .
Constants in the estimates
Recall the constants from (3.14). In later proof, we will use the following constants, which have size of compared to and
(3.26a) | ||||
(3.26b) |
3.3. Consequence of the inductive hypothesis
Suppose that the inductive hypothesis holds true for the case of and . We show that for with
(3.27) |
where are chosen in (3.23), we have the inequality 777 We use instead of in the upper bound since would be .
(3.28a) | ||||
The estimate is trivial if or . Due to (3.27), it remains to consider . Since (3.28a) is symmetric in and , we may without loss of generality assume | ||||
(3.28b) |
Since , , from the choices of in (3.23), we get
(3.28c) |
To prove (3.28), we bound from below and from above. Since due to (3.23) and , using (3.22e) repeatedly, we get
(3.29) |
where in the last inequality, we have used since
for . Since we consider , can be negative.
The parameter can be seen as the accumulated error between the actual ratio and the desired asymptotics in (3.22e) with much larger than . Next, we estimate from above and consider and .
Case 1:
Since due to (3.23) and (3.27), we get
Each product is non-negative since (3.28b), (3.27). Moreover, since (3.23) and for , we obtain
We want to show that . Since (3.27) and in this case, using , we obtain
Combining the above estimates, we derive
Due to the choices of in (3.23), for any , we have . Since and , we obtain that is increasing in . Thus, we obtain
(3.32) |
Case 2:
3.4. Estimate of
We will use (3.16) to estimate the coefficients . Below, we estimate . We introduce the truncated power series of up to
(3.33) |
Since we choose in (3.23) and fix in the following estimates, to simplify the notation, we drop the dependence of on . Note that the above power series involves rather than . We recall the relation between from (3.20).
Denote by the coefficient of in the power series expansion of . Next, we estimate for (2.15a) and .
For and , we get
(3.34) |
We use the inductive hypothesis to obtain two estimates of with . The first estimate applies to and the second applies to .
First estimate with .
Without loss of generality, we assume . Since and , we get . Using (3.22a) repeatedly, we get
(3.35) |
Due to , the constraint on (3.34), and by the choices of (3.23), we obtain that satisfies
(3.36) |
It is easy to verify that the conditions (3.27) hold for . Thus, using (3.28) with and (3.35), we yield
Thus the above estimate imply
(3.37a) | |||
where is defined in (3.26). |
Second estimate with .
Using (3.22a) and , we get another estimate
(3.37b) |
Using the identity for the Catalan number (3.20b) repeatedly, we get
(3.37c) |
For , we just use the definition (3.33) and (3.20) to obtain the th coefficient of : . Since we define (3.26a), the above estimates hold trivially for .
Remark 3.8.
3.5. Proof of the induction
We decompose the right side (RS) of (3.16) into three parts as follows: the first term, the summation over with and with
(3.41) | ||||
We estimate in Sections 3.5.1, 3.5.2, and the lower order terms on the left side of (3.16) in Section 3.5.3. Our goal is to show that they are small compared to , e.g.,
We combine these estimates to prove the induction in Section 3.5.4.
3.5.1. Estimate of
Recall from (2.15b). Expanding the coefficients of using (3.2), we get
(3.42) |
where denotes the -th order coefficient of in the power series of . For , since and from the definitions (3.23) of and (2.15c) for , we get
Therefore, we may restrict the summation to in (3.42). Next, we estimate
(3.43) |
and consider with and separately. We have due to the constraint in (3.42). Since we fix below, we drop the dependence on .
Case 1:
In this case, using (3.40), we get
In this case, since and ( (3.23)), we can bound by some constant (3.26) independent of and obtain
Denote
From the bound of , we get and can further estimate as follows
(3.44) | ||||
The bound in the maximum is very small by choosing relatively large and then large.
Case 2: .
If , using (3.28) with , , and again, we obtain
(3.45b) | ||||
If , using (3.28) with , similarly, we get
Since (see (2.15c)), satisfies and , we obtain
Since satisfies , we use (3.22c) to obtain . Using the above three estimates, we simplify the bound as
(3.45c) |
Combining the estimates in (3.45), we establish
(3.46) |
Summary
Summing two estimates of (3.44), (3.46) for two cases, we establish
(3.47) |
for any , where is defined as
(3.48) | ||||
and are defined in (3.26). The first term in corresponds to the bound (3.44) for Case 1, and the second term for the bound (3.46) in Case 2.
Since defined in (3.26b) is of order , by first choosing large so that the second part in is small and then large 888 Choosing large means that we verify the estimates in Lemma 3.5 with computer assistance up to the case with large . so that the first part is small, we can make very small. Therefore, combining the above estimates, we can estimate as follows
Summing over using (3.20b) for , we yield
(3.49a) | |||
Using (3.20c) for any , we further bound and obtain | |||
(3.49b) |
Since can be made sufficiently small by choosing in order, can be made very small.
3.5.2. Estimate of
3.5.3. Estimate the LHS of (3.16)
Below, we estimate the left side (LS) of (3.16). We decompose the left side as follows and treat the terms of perturbatively
(3.53) |
Using the estimate (3.39) for , we estimate the error term as follows
Below, we further estimate and . For , using (3.14) we get
Since satisfies and (3.23), using (3.20), we obtain . Combining these estimates for and denoting , we conclude
(3.54a) | ||||
Using , we can rewrite the above summation and obtain the following estimates | ||||
(3.54b) |
For fixed , since , is of order .
3.5.4. Proof of the induction
Now, we are in a position to prove the induction and Lemma 3.5, which is based on the following lemma verified with computer assistance.
Lemma 3.9 (Computer-assisted).
Proof.
Estimate (3.22e). Recall (3.14b). Combining the estimates (3.41), (3.49), (3.51), (3.52) of for the right side of (3.16) and the estimates (3.54), (3.53) for the left side of (3.16), we can estimate the relation between as follows
Using continuity of the functions in and Lemma 3.9, for sufficiently large, we obtain
From Lemma 3.1 and (2.30), we obtain . Using (3.20) and diving the above estimate by , we compute
Using triangle inequality, we obtain
From (3.20), we have . For , using (3.55c), the limit (3.22d) and continuity of in , we obtain
Thus, for sufficiently large (equivalent to close to ), we establish and (3.22e).
Estimates (3.22a) and (3.22c). Using the new asymptotics for and defined in (3.26b) and for , we obtain . Since from (3.55), we verify in (3.22c). Combining this estimate, (3.28) with , and using from Lemma 3.9, we establish
for any . Since from item (a) in Lemma 3.6, we verify (3.22a).
We prove all the estimates for in the induction and thus establish Lemma 3.5.
3.6. Convergence of power series
In this section, we establish the following estimates of the power series coefficients and establish the convergence of the power series.
Proposition 3.10.
Recall the definitions of from (2.33), (2.34). Let with defined in (2.15a), be a positive integer, and be a closed interval. There exists a constant depending on such that the renormalized power series coefficient (3.20) with parameter satisfies
(3.56) |
In particular, for any , the solution with defined in (2.33) and is analytic in . Moreover, is continuous in and .
Here, we use the notation rather than to indicate the dependence on the parameter (or equivalently ). This convention will simplify our notations in Sections 4, 5 when referring to this local analytic solution.
Proof.
Firstly, for any , using (3.17) and (2.30), we obtain
For with , from the definition of (2.34), we obtain for a close interval . Thus, we obtain with some uniformly for any and . From (2.28), (2.27), we obtain . Combining these estimates, we obtain
(3.57) |
uniformly for . In particular, is bounded uniformly for .
Next, we prove (3.56) by induction. We fix an arbitrary and drop the dependence of on to simplify the notation. For to be determined, we choose large so that (3.56) holds for . Suppose that (3.56) holds for the case of with . We use (3.17) to bound . We follow the arguments in Sections 3.4, 3.5 with .
Following (3.33) with , we introduce the truncated power series For and , the power series coefficient of is given by (3.34)
(3.58) |
Using the inductive hypothesis (3.56), we get
(3.59) |
Without loss of generality, we assume and for some with . If and , using the definition of in Proposition 3.10 and , we get
If and , we obtain
If , we obtain and
Plugging the above estimates in (3.59), applying the identities (3.20b) repeatedly, we obtain
(3.60) |
for and . Since and , (3.60) also hold for .
Next, we bound the right hand side of (3.17). Following Sections 3.5, we only need to bound in (3.41) with . For (3.41), using the expansion (3.42) with , we get
Applying (3.60) for , inductive hypothesis (3.56) for , we obtain
Since , and is large, it is easy to get that the exponent is bounded by . Using (3.20), we get
For (3.41), we bound it similarly
Continuity
Lastly, we prove the continuity of in , where
and . Fix an arbitrary . We decompose
where is the inverse map defined in (2.33). From the above estimate of , by choosing large, we obtain uniformly in . From (3.57) and (3.17), each term with is uniformly continuous in . Since the map is uniformly continuous in for , we obtain that is uniformly continuous in . In particular, there exists small such that for any .
3.7. Refined asymptotics
As a consequence of (3.22e) in Lemma 3.5 and (3.20), we have the following estimates of the original coefficients .
Corollary 3.11.
Proof.
Using (3.20), we decompose
From the asymptotics (3.20b), we obtain . Since from (3.22e), and (2.10), taking sufficiently large so that is sufficiently close to , we prove for any .
Using Lemma 3.5 and Corollary 3.11, we have the following refined estimates of , which will be crucially used to estimate barrier functions.
Lemma 3.12.
Let be the parameters defined in (2.10) and Corollary 3.11. There exists large enough, such that for any with , the following statement holds.
For any with and and , we have
(3.62) |
For any with , we have
(3.63a) | |||
(3.63b) | |||
(3.63c) |
For any and with , then for
(3.64a) | |||
we have the estimate | |||
(3.64b) |
We have
(3.65a) | |||||
(3.65b) | |||||
(3.65c) |
Estimate (3.62) is a refinement of (3.22a). In (3.63), we compare with the reference scale and show that the former is much smaller if is away from or , and it is not much larger if is close to . Estimate (3.63c) will be useful for with being some absolute constants. Such an estimate is not covered by (3.63a), (3.63b) as we cannot choose comparable to in (3.63a), (3.63b) (otherwise (3.63a), (3.63b) are trivial). In (3.64), we show that grows exponentially to estimate the power series of in later proof of Proposition 5.1. We will only apply (3.64) with away from . In (3.65), we show that is very large relative to and . In the following proof, we will treat the parameters (3.23), e.g. in Corollary 3.11 and the bound of as absolute constants.
Proof.
Since (3.62), using a change of variables , and combining the above two estimates, we obtain
We estimate the product. Since , , , and , we obtain
Proof of (3.63b). We only need to consider . In this case, for all , we have . Since (2.10), using the above estimate, we yield
Proof of (3.63c). Choosing in the above estimate (3.67) and using , and for , we further obtain
(3.68) |
Next, we consider . Denote . From the assumption of above (3.64), we obtain . By requiring large, since , we also have . Since , using (3.61), we get
As a result, we yield
(3.69a) |
Since (3.61) for , using the above estimate again, we get
(3.69b) | ||||
Proof of (3.65c). Next, we prove (3.65c), (3.65a), and (3.65b) in order. Recall chosen in Corollary 3.11. Using (3.61) and for , we prove
Proof of (3.65a). If , the estimate (3.65a) follows from (3.65c). Next, we consider . Using estimates similar to (3.66) and for , we get
(3.70) |
Since , combining (3.70) and the upper bound of (3.68), we estimate
The common factor in the denominator and numerator is cancelled.
4. Smooth solution connecting and
In this section, we study the phase portrait of above the point , which corresponds to the region in the original ODE (2.3). Our goal is to prove the following result.
Proposition 4.1.
There exists large enough such that for any with , there exists and , such that the following statement holds true. The ODE (2.3) admits a solution with , and
(4.1) |
for any , and for some function . Moreover, it agrees with the local analytic function near with constructed in Proposition 3.10 in the following sense
(4.2) |
for with small , where are the maps defined in (2.12).
Throughout this section, we use to denote an integer rather than a dummy index and will first consider or . In Section 4.4, we will consider and prove Proposition 4.1 using a shooting argument.
Ideas and barrier argument
We first construct the far-field lower and upper barriers with (see (4.5b) and (4.5a) for the definitions). If is outside the region
(4.3) |
the solution will remain outside for any (see Section 4.1 and Proposition 4.2). We define the boundaries of
(4.4) |
See Figure 2 for illustrations of , and the solution curve.
We construct the local upper and lower barriers based on (4.14).
In Propositions 4.3, 4.4, we will show that is a upper barrier for , which is valid for . Then, in Propositions 4.5, 4.6, we show that intersects at some for some with large and small . This implies that the solution exits the region via for this . We perform a similar argument for the lower barriers and some with large and small . See Sections 4.2, 4.3.
In Section 4.4, using continuity, we construct a smooth solution from to with close to , and the solution agrees with the local analytic solution to (2.3) near via the map (2.11). Using the inverse map (2.12) proves Proposition 4.1.
Here, u, l, ne, f are short for upper, lower, near(for local), far, respectively.

4.1. Far-field barriers
We construct the far-field upper barrier
(4.5a) | ||||
and the far-field lower barrier | ||||
(4.5b) | ||||
with satisfying | ||||
(4.5c) | ||||
(4.5d) |
For (4.5c), a direct calculation obtains the derivative and solves :
We impose (4.5d) to show that the quantity in the bracket in (4.5d) is negative. See (4.7d). It is easy to see that (4.5d) is linear in , and we can solve easily with symbolic computation.
Recall the roots from (2.24), and from (2.23). In Propositions 4.2, 4.4, we will show the following relative positions of curves near
(4.6) |
for . See Figure 2 for illustrations of the relative positions of these curves.
We have the following results.
Proposition 4.2 (Computer-assisted).
There exists a large such that for any , the following statements hold true. The functions (4.5) satisfies and
(4.7a) | ||||
(4.7b) | ||||
(4.7c) | ||||
(4.7d) | ||||
(4.7e) |
where . Moreover, the function (2.23) and (2.24) satisfies
(4.8) |
for some constant independent of . The map (2.14) along the solution curve satisfies
(4.9) |
As a result, we have for defined in Lemma 2.2, and
(4.10) | |||||
(4.11) |
and are an upper, and lower barrier for the ODE with , respectively.
The proof involves computer assistance, and we refer to Appendix B for further details.
Proof.
For (4.9), since is smooth near and
using chain rule, we prove the first identity in (4.9). The second term in (4.9) is a scalar and only depends on . We do not expand it below as we can derive it symbolically.
Methods. Recall that choosing sufficiently large is equivalent to choosing close to (2.31), (2.32). Each inequality in (4.7) and (4.8) can be reformulated equivalently as
(4.12) |
for some , where are polynomials in and continuous in . We verify that
uniformly with some using computer assistance. See Appendix B for more details. Then using continuity, we prove (4.7) for close to .
Consequences. Recall from (4.3) and from Lemma 2.2. Since (2.16) and (4.8), we obtain that (4.5a) is increasing for . Therefore, for any (4.3), we obtain , and .
Estimate (4.10) Since both (2.23) and (4.5a) are quadratic polynomials with , we get
Since (2.1), and from (4.8), we prove for any and obtain the first estimate in (4.10).
To prove (4.10), using the formula (2.22), we only need to show
(4.13) |
For any with defined in Lemma 2.2, using the first estimate in (4.10) and Lemma 2.2, we have and , which implies (4.13)
4.2. Local barrier
For with large to be chosen, we construct the local barrier as
(4.14) |
We define
(4.15) |
We have the following results regarding the sign of .
Proposition 4.3.
There exist large absolute constants such that for any , and , the following holds true.
(a) For any and , we have ,
(b) For any and , we have .
Here, is some constant depending on .
Proof.
We require larger than the constant in Lemma 3.12. Since , we first choose small enough so that
(4.16) |
In the following estimates, we will mainly track the dependence of the constants on , which plays a role as a small parameter, and , which are large.
Below, we shall simplify as , and use to denote the -th coefficient in the power series expansion of . We will show that
(4.17a) | |||
where is given by | |||
(4.17b) | |||
and satisfies the bound | |||
(4.17c) |
From (4.14), (4.15), is a polynomial in . Since the barrier (4.14) and the power series of agree up to , using Lemma 3.2, we get that the term in power series expansion of vanishes. Moreover, using the notation (3.11), and the derivations (3.10), (3.15) with and from (3.17), we obtain the coefficient of in
which is (4.17b).
Next, we estimate the coefficients of with . Applying Corollary 3.11 and tracking the dependence on , we obtain
(4.18) |
Next, we estimate the coefficients of the power series of . Since is linear in (2.14), using (4.18), we obtain
Since is quadratic in (2.14) with the nonlinear term , using the product formula (3.2) and the estimates (4.18), yields
Recall from (4.14). Using
the above estimates of of , and the product formula (3.2), for , we obtain
and establish (4.17c).
Next, we estimate and . Using (2.30), we first rewrite as follows
Choosing with and using (4.18), we get . Since (2.28), , and , we estimate
To ensure that the error term in (4.17) is smaller than , e.g. , we require
which are achieved by choosing sufficiently small in (4.16). As a result, we get
Recall that we assume or . Since from (3.22b), using the relation between in (3.61), we obtain
which implies
We conclude the proof.
Next, we derive the relative positions among the barriers and the solution.
Proposition 4.4.
For , we will use as the local upper barrier for , and (Y) as the local lower barrier.
4.3. Intersection between local and far-field barriers
In this section, we estimate the location where the local and far-field barriers intersect, and then show that the intersection occurs within the region where the local barriers are valid.
Proposition 4.5.
Let be the parameters chosen in Proposition 4.3. For any with , where is some large parameter, and satisfying the assumption in Proposition 4.3, there exists such that the following statements hold true.
-
(a)
For any , there exists satisfying
such that
(4.23) -
(b)
For any , there exists with
such that
(4.24)
Proof.
By definitions of (4.5), (4.14), we have
Hence, these functions agree at up to . Moreover, from Proposition 4.4, we have
Below, we focus on the proof of (4.23). For , from (3.61), (3.22b), we get
(4.25) |
Using the assumptions of in Proposition 4.3, for , we obtain
which along with (4.25) imply
(4.26) | ||||
Using this estimate, (4.19a), and continuity, we obtain that and intersect at some . We assume that is the intersection with the smallest value in . The smallest of and (4.19a) imply the first inequality in (4.23). Moreover, we have
(4.27) |
For a fixed , by choosing small enough, we can ensure that . Moreover, since , from (4.7a), we obtain.
Thus, by further requiring small, which leads to a smaller upper bound for in (4.27), and using (4.26), we establish
for . Rearranging the inequality, we prove the second inequality in (4.23) and the result (a) in Proposition 4.5.
The result (b) is proved similarly using the property that in such a case.
Proposition 4.6.
For any with chosen in Proposition 4.5, we have the following results. For , there exists , such that the local smooth solution starting at with parameter first intersects at .
Proof.
From Proposition 4.4 and the property that (4.14) agree at up to error , we know that
remain in (4.3) for with small .
We focus on the proof of the case with . We fix and choose with
(4.28) |
where are the parameters chosen in Proposition 4.3, and in Proposition 4.4, respectively. From Proposition 4.4 and the above discussion, we have
(4.29) |
for with sufficiently small. Since in from Proposition 4.4, and for and with
using a barrier argument, we obtain that
(4.30) |
Next, we choose in the range of defined in Proposition 4.5. From the choice of (4.28) and the inequality of in Proposition 4.5, we get
Due to (4.29) and the fact that intersects for within the validity of barrier (4.30), the solution must intersect at some . Using (4.23) in Proposition 4.5, we get
Thus, cannot intersect for . We conclude the proof in the case of .
The proof of the case of is similar and is omitted.
4.4. Proof of Proposition 4.1
In this section, we first construct the smooth solution to the ODE (2.3) near . Then we use a shooting argument to glue the curve under the map (2.11) and the smooth solution starting from and prove Proposition 4.1.
We have the following result from [56, Proposition 3.3] and its proof.
Proposition 4.7 (Proposition 3.3, [56]).
Let be the coordinate of the sonic point in (2.7). The ODE (2.3) has a unique solution with . Moreover, near , it has a power series expansion 999 In [56], the authors rewrote the ODE (2.3) as an ODE for and and expand as a power series of . In particular, the smooth local solution can be written as for some smooth functions .
(4.31) |
where is the Catalan number (3.20) and is some absolute constant. 101010 Although, the constant depends on the parameters , since we restrict (2.10) to specific ranges, following the same estimates as those in [56] lead to an absolute constant.
We only need the local existence of for with some . The proof is standard and also follows from the argument in [6, Section 2] by bounding the power series coefficients. The power series of only contains the odd power due to symmetry.
Using the map (2.11) from to coordinates and Proposition 4.7, we construct
(4.32) |
where is short for far. See the orange curve in Figure 2 for an illustration of .
We have the following asymptotics of for small .
Lemma 4.8.
The proof follows from expanding near and using the map (2.11), which is elementary but tedious. We defer it to Appendix A.2.
Lemma 4.9 (Computer-assisted).
We have
We obtain the limit by definition of (4.5b). We verify the scalar inequality in Lemma 4.9 directly using Interval arithmetic.
Since and is uniformly bounded for near , using Lemma 4.9 and Lemma 4.8, we obtain that there exists with chosen in Lemma 4.8 such that
(4.35) |
for any and any .
4.4.1. Shooting argument
Denote by the local analytic function near constructed in Proposition 3.10. We prove Proposition 4.1 using a shooting argument. Let be the large parameter determined in Propositions 4.5. We fix and consider .
Recall the region from (4.3). From (4.7c), in Proposition 4.4, for any , we have for . Moreover, since in from Proposition 4.4, the solution curve either remains in for all or exits the region via one of the edges (4.4). We define the extension of according to one of two cases.
(a) If for , we define
(4.36a) |
(b) Otherwise, suppose that first intersects at with for . We define
(4.36b) |
with if and if . That is, we extend using the barrier function or .
Since is in the closure of , where we have if , using the continuity of the ODE solution to (2.14) in , we get that is continuous in .
From Proposition 4.6, for , there exists and such that:
-
i)
The solution first exit at with .
-
ii)
The solution first exit at with .
Since , using the estimate of (4.33a), we can choose small enough and independent of such that
(4.37) |
We define
4.4.2. Gluing the solution
We construct solution to the ODE (2.3) using the maps (2.12) from and then glue it with obtained above.
We use the map (2.12) to construct the curve
(4.39) |
for some . Using (4.39) with and (4.38), we get
(4.40) |
Using the second identity in (2.19) and the chain rule, we obtain
(4.41) |
Next, we show that is strictly decreasing and invertible. Using (2.19), we get
(4.42) | ||||
Recall, in Section 4.4.1, we showed for ; hence, from (4.11) and (4.10) we obtain the inequalities . Using these sign inequalities, together with (2.19), yields
Using (4.9) and the definition of (4.39), we get . Using continuity, we obtain
with and . Thus, is strictly decreasing, invertible, and smooth on . We define the inverse as and construct a solution
(4.43) |
Using (4.43), (4.41), and the chain rule, we obtain a smooth solution to the ODE (2.3). Using (4.40) and (4.38), we get
4.4.3. Proof of other properties
Since map the sonic point to in the system, using (4.39) with and (4.43), we get . Since for small is constructed by Proposition 4.7, we have and for some with small .
To prove (4.1), using Lemma 2.2 and the property that (2.11) with is equivalent to
we only need to estimate the coordinate of the solution curve:
(4.44) |
with some . From Lemma 4.8 and the definitions of (4.3) and (4.5a), we have
for and . From the relation (4.32) and (4.39) with and (4.40), we prove (4.44) for . Using continuity and by choosing small enough, we prove (4.44) for . We conclude the proof of Proposition 4.1.
5. Lower part of
In this section, we consider odd and any . We use a barrier argument to show that the local analytic solution constructed in Proposition 3.10 can be continued for and cross the curve (red curve) below the point (see Figure 2). We justify these in Propositions 5.1, 5.2, 5.3. Since the curve below in the system (2.15b) is a upper barrier and is another barrier, the solution curve must further cross (2.11) (see Figure 2 for an illustration), which corresponds to the curve of in the original system (2.3) (see the red curve between in Figure 1). We justify it in Proposition 5.6. Afterward, in Lemma 6.2 in Section 6, we can extend the solution of the (2.3) to . This give rises to a smooth solution to the ODE (2.3) for for some .
5.1. The local upper barrier
We introduce a local barrier
(5.1) |
where are the power series coefficients constructed in Section 3. In Proposition 5.1, we show that remains a valid local upper barrier of for , where is defined in (3.22d). In Propositions 5.2, 5.3 we show that in the range , the local upper barrier and the solution must intersect the global upper barrier .
Proposition 5.1.
Proof.
Recall from (2.10). For to be chosen, we define according to
(5.3a) | ||||
Since and , by choosing close to , we can obtain satisfying
(5.3b) |
Recall from (3.22d). We define
(5.4) |
We estimate for separately.
Expansion of
Denote by the -th coefficient in the power series of . Using the formula and the definition of (5.1), we obtain
(5.5) |
Next, we perform power series expansion for (5.2). Denote by the coefficient of in . Since agree with the local solution up to , following the derivation of in (4.17b) in the proof of Proposition 4.3, we obtain that the term vanishes in and the leading order term of is given by
(5.6a) |
For the coefficient of , using the explicit formula of (2.14), we obtain
(5.6b) | ||||
We note that the term in and in (2.14) only contributes to the term in .
Thus, for odd and , we get .
Ideas of the remaining estimates
It remains to estimate the term for . We use the asymptotics for in Lemma 3.12 to show that . For very small, we treat all the terms as perturbation to . For relatively large and close to , we treat as perturbation to . We choose to be some large absolute constant, and perform the estimates in four cases:
Case I:
Using (3.61), we obtain
Thus, for , using the above estimate and the lower bound of (5.8b), we obtain
(5.10) |
By choosing , we can treat it as perturbation to .
Case II:
Case III:
Denote
(5.13) |
In this case, we have . Using (3.63c) with chosen in (5.3) and (5.11b), for with (5.4), we obtain
(5.14) | ||||
Next, we further simplify the upper bound. We introduce
We estimate in the case of and separately. Using (5.3), we get
Denote . From (5.13) and the assumption of , we get . Using and in (5.3), we further estimate the second case as follows
Combining the above estimates, we establish
(5.15) |
Case IV:
Recall from (5.4). We consider and separately. We define as (5.13). For , following (5.11) and applying (3.63a) with , we obtain
(5.16) | ||||
Since (5.3b), the upper bound is very small compared to .
Combining (5.10), (5.12), (5.15), (5.9), and (5.16), summing these estimates over , and using the lower bound of (5.8b), we prove
(5.17) |
where we use to denote .
Next, for and with , we show that
Since and , using (3.64), we get
Using (5.3) and (5.4), we yield
Since (5.3), for sufficiently large and , since , we prove
(5.18) |
We conclude the proof.
5.2. Intersection between the barrier functions
Next, we estimate the first intersection between the local barrier and the global barrier. Recall defined in (2.24).
Proposition 5.2.
Let be the parameter defined in Lemma 3.12. There exists large enough, such that the following statement holds true. For any , there exists with such that
As a result, we have
Proof.
Denote
(5.19) |
Firstly, using a direct computation, we know that
Using the above estimate and the definition of (2.24), for and sufficiently small, we obtain
(5.20) |
Since is smooth near , for some absolute constant and , we have
(5.21) |
It follows
Choosing large and then large, we further obtain
Since is odd, it follows
Since and in (5.21), (5.23) is absolute constant, by further requiring large enough, we yield . Thus using (5.23), we obtain
Since for (5.20) and is continuous, there exists with such that and for . Using (3.65), since (2.10) and , by further choosing large, we obtain
(5.24) |
Below, we show that the solution must intersect the curve (2.24).
Proposition 5.3.
For any odd large enough, we have the following results. Let be defined in Proposition 5.2. There exists and such that and
(5.25) |
Proof.
Recall the definition of from (5.1). Firstly, we have for and
Since is odd, using and (3.22b), (3.61), we obtain . Using the above estimates and
we establish
for with some small parameter .
Since intersects at , by continuity, we obtain that intersects at some .
Since from Proposition 5.2 and , by choosing large enough, we obtain . We complete the proof.
We have the following relative positions among the roots and the barrier functions.
Proposition 5.4.
Proof.
We consider . Denote . From the definitions of (2.9), (2.24), and (5.1), we have . Thus, we only need to show
for and some absolute constant independent of , where or . Since , using (4.8), we only need to show that the inequality
(5.28) |
with some absolute constant independent of and determined in (4.8) holds for any . By requiring large, we have (5.24) and
Since is odd and , using (3.22b), we get and . Below, we further bound .
For , using (3.63b) with and , for , we get
(5.29) |
For , applying (3.65a) and from (3.63a), we obtain
(5.30) |
For , we estimate the cases and separately.
For , using (3.63a) with and , we get
(5.31a) | |||
Combining (5.29)-(5.31a) and , we prove (5.28) for large enough with . |
We have the following estimate of and near .
Lemma 5.5.
There exists constant such that
for any . Moreover, we have for any .
5.3. Extension across
In this section, we show in Proposition 5.6 that after crosses , we can smoothly extend the solution curve slightly beyond . See Figure 2 for an illustration. In Section 6, we will use Proposition 5.6 to prove Theorem 1.1.
We introduce the triangle region below
(5.32) | ||||
From the definition of in (5.32) and (2.24), we get
(5.33) | ||||
Since the ODE becomes singular on , we desingularize the ODE (2.14)
(5.34a) | |||
where ds is short for desingularized. Let be the intersection determined in Proposition 5.3. Using estimates of in (5.25),(5.27), we can pick the initial data as | |||
(5.34b) |
Define
(5.35) |
Since for , we define a map
(5.36) |
We have and it is a bijection. For the desingularized ODE, we show:
Proposition 5.6.
There exists with such that:
(a) The solution can be obtained from by a reparametrization.
(5.37) |
(b) For any , we have
(c) We have for any and ;
(d) For any , we have . For any (including ), we have
(e) In particular, for any , we have
5.3.1. Estimates of
We define . Since is integrable near from Lemma 5.5, using the equation of (5.36), we obtain that is bounded.
Using the ODE (2.14), the definition of (5.36), and the chain rule, we obtain that the ODE system of in is the same as (5.34). Moreover, two ODEs have the same initial condition at due to (see (5.36)) and (5.34b). Using the uniqueness of the ODE solution and continuity, we prove (5.37).
5.3.2. Solving the desingularized ODE
Next, we solve the ODE (5.34) for and prove item (c), (d), (e) in Proposition 5.6. We consider
Using (5.38), we get for and . Moreover, using the ODE (5.34), we compute
We obtain the last inequality using the strict inequalities in result (a) in Proposition 5.6 at and (2.24). Using continuity, for with small , we obtain
(5.39) |
Intersect
Next, we show that intersects the curve . Due to the sign conditions (5.33), starting at , we know that if , we have
(5.40) |
Due to continuity and (5.39), we get for with some . We assume that is maximal value such that for all and is allowed to be . Below, we show that and .
For , since and is increasing in for from Lemma 2.4, using (5.40), we get
(5.42) |
for some constant independent of , where the last inequality follows from (5.32). Using (5.41), (5.42), and the uniform boundedness of (5.41), we estimate in (2.14) as
(5.43a) | ||||
(5.43b) |
for all and some constant independent of .
Due to (5.43b) and for any , the curve must exit with finite . Using (5.43a) and the ODE of in (5.34), we get . Since are continuous in , taking in (5.42), we get
(5.44) |
Using (5.44), the definition of in (5.32), and , we obtain that can only exit via . Thus, we prove and the result (c) in Proposition 5.6.
Crossing the line
Recall that (5.43b) applies to any . Since are continuous, taking , we get
6. Proof of Theorem 1.1
In this section, we prove Theorem 1.1.
Recall from (2.20), which relate to the roots of . We introduce
(6.1) |
The proof of Theorem 1.1 follows from the following two results.
Proposition 6.1.
Lemma 6.2.
Given any with and , the ODE solution to (2.3) starting at admits a smooth solution for any with and for some .
We defer the proofs of Proposition 6.1 and Lemma 6.2 to Sections 6.1 and 6.2, respectively. See the black curve in Figure 1 for an illustration of the solution curve . The region lies to the right of the red and blue curves in Figure 1.
Proof of Theorem 1.1.
Using Proposition 6.1, there exists large enough, such that for any , is odd, and some , we can construct a smooth solution to the ODE (2.3) in with the properties in Proposition 6.1. Since and (6.1) is open, we can choose with . Applying Lemma 6.2 with , and using the uniqueness of ODE, we construct a global smooth solution to the ODE (2.3) with and the estimates (1.9).
From Proposition 4.1, we obtain for some with some . Since , we can extend this property for some .
From (1.9) and Lemma 6.2, we obtain , ,
(6.3) |
Using the power series expansion (4.31) in Proposition 4.7 and (4.34), we have . Thus, using the above estimates, we obtain that the denominators in are non-zero for , the singularity is cancelled near , and .
For any large enough, estimating using the ODE (2.3), we derive the asymptotics
Choosing , we obtain
(6.4) |
with . We further construct the profile using (A.11) with . Using these profiles, the self-similar ansatz (2.2), (1.3), and (1.7a), we construct the smooth profile for and obtain a smooth self-similar imploding solution (1.10) to the relativistic Euler equations (1.7). The asymptotics (1.11) follows from the limits of in (6.3), (6.4) and the formulas of in (1.7a), (1.10),(2.2). We complete the proof.
6.1. Proof of Proposition 6.1
We assume that is odd and large enough. Using Proposition 4.1, we construct a smooth solution with some and
(6.5) |
First gluing
In Proposition 5.3, we show that can be extended smoothly for up to . We define the map following (4.39)
From Proposition 5.3, (5.27), we have for . Thus, along the solution curve with , using (2.24), (2.22) and following the proof of (4.10), we obtain . Following (4.42), we obtain
for any . Using and continuity, we obtain
(6.6) |
and is invertible for . Constructing a solution to the ODE (2.3), and using the gluing argument in Section 4.4.2, we extend the relation (6.5) to :
(6.7) |
Moreover, solves the ODE (2.3) smoothly for .
Second gluing
Next, we construct another solution using the desingularized ODE (5.34)
(6.8) |
where is defined in Proposition 5.6. Below, we consider arbitrary . We show that . Using the ODE (5.34) and the second identity in (2.19), we yield
From result (e) in Proposition 5.6, we have . Using (2.22), we obtain . Since , due to (2.19), we obtain . Using (6.6) and continuity, we obtain and . Thus, is invertible.
Estimates (6.2)
Recall that is the sonic point (2.7). We define
(6.10) |
From Propositions 5.3, 5.6, and Lemma 5.5, the solution curves with and with are in . From (6.7), (6.8), under the map (2.12), these curves are mapped to with . Following the proof of (4.1) in Section 4.4.3, we prove the estimates (6.2) with . Estimates (6.2) with have been established in Proposition 4.1 .
Location
6.2. Proof of Lemma 6.2
By continuity, there exists such that the solution to the ODE (2.3) exists and for . We assume that is the maximal value with such a property. Our goal is to show that . Since do not vanish for , and are polynomials in , we get .
Since is increasing in (2.21) and , we obtain
(6.11b) |
Since for all , it follows
(6.11c) |
for some constant . For with , since , we estimate (2.3):
(6.11d) |
for some constant . For any with , combining (6.11), we have
which implies
(6.12) |
If is bounded, we choose . For , since is bounded, is bounded (6.11d), and is bounded away from (6.11c), we can solve the ODE (2.3) with solution in with sufficiently small. Using the uniform estimates (6.11b),(6.12) and choosing small enough, we obtain that (6.1) for . This contradicts the maximality of . Thus, we have .
Appendix A Some derivations and estimates
A.1. Equations of the profiles
The ODE (2.3) has been first derived in [56]. For completeness, we derive it below. With the ODE solution, we can further construct the profiles in (1.3).
We consider radially symmetric solutions to (1.2b) and (1.5). Recall the self-similar ansatz (1.3) and the notations (2.2)
(A.1) |
We will only use the notations in this section.
Since and commute for smooth functions, we get
(A.2) |
Using the self-similar ansatz (1.3) with , we get
(A.3) |
We further compute . Dividing on both sides of (A.2) and using (A.3), we get
which implies
(A.4) |
Next, we use (1.2b) to derive the equation for . Since is radially symmetric, using , we can rewrite (1.2b) as
Using (A.1) and , we further rewrite it as
Dividing on both sides and using (A.3), we obtain
(A.5a) | ||||
(A.5b) |
Since , for , we obtain
(A.6) |
Using the formulas of (A.1), we compute
(A.8) |
Rewriting the above equations, we derive the ODE (2.3):
Once we construct the ODE solution with , we further construct the profile (1.3) for . Plugging (1.6), (A.6), (A.8), , in (A.4), and cancelling the power of , we obtain
(A.10a) |
Using (A.9) , we can rewrite the right hand side as
(A.10b) |
A.2. Proof of Lemma 4.8
Using Proposition 4.7, we can expand the smooth solution near as follows
(A.12) |
for . In the above and the following derivations, the implicit constants and are uniformly in due to Proposition 4.7.
Firstly, we compute . Using the ODE (2.3) and evaluating at , we obtain
(A.13) |
Next, we compute . We expand near up to the term
Using the ODE (2.3) and tracking the term up to , we get
Matching the term yields . Matching term, we get
Rearranging the identity, we obtain in (4.34). Identity (A.12) gives the formula of in (4.34).
Appendix B Details of the computer-assisted part
In this section, we discuss the companion files for computer-assisted proof. We performed the rigorous computation using SageMath, and the code is attached to the paper. There are two Sage files: Wave_induction.ipynb, Wave_barrier.ipynb, which can be implemented on a personal laptop in less than 1 minute. Moreover, all the outputs are recorded in a Jupyter Notebook without implementing the codes.
For the variables/functions with different notations in this paper (left side) and in the code (right side), we provide the relation between two notations below:
The functions and parameters are defined in: (2.9), (2.6), (2.14), (2.10), (2.27a), (2.24), (3.3), (3.5).
Wave_induction.ipynb
(1) Derive the power series coefficients at the limit case using the recursive formula (3.17) in Lemma 3.2.
(3) Verify Lemma 3.6 using the value at the limit case derived in Step (1).
Wave_barrier.ipynb
The goal is to verify (4.7) and (4.8) in Proposition 4.2 and the scalar inequality in Lemma 4.9. It consists of the following steps.
(1) Derive the power series coefficients at the limit case , which is the same as those in Wave_induction.ipynb. We only need .
(2) Introduce a function Poly_sign to derive upper and lower bounds of a polynomial over with . We can decompose with being polynomials with non-negative coefficients. For , is increasing. Thus, by dividing into and using
for , where
we estimate the upper and lower bound of in each interval . Maximize or minimize the estimates over all sub-intervals yields the bounds of in .
Acknowledgments
The work of T.B. has been supported in part by the NSF grants DMS-2243205 and DMS-2244879, as well as the Simons Foundation Mathematical and Physical Sciences collaborative grant ’Wave Turbulence.’ The work of J.C. has been supported in part by the NSF grant DMS-2408098. We would like to acknowledge the insightful and fruitful discussions we had with Hans Lindblad and Jalal Shatah.
References
- [1] Leo Abbrescia and Jared Speck. The emergence of the singular boundary from the crease in compressible euler flow. arXiv preprint arXiv:2207.07107, 2022.
- [2] Leonardo Abbrescia and Jared Speck. The relativistic euler equations: Esi notes on their geo-analytic structures and implications for shocks in 1d and multi-dimensions. Classical and Quantum Gravity, 40(24):243001, 2023.
- [3] S. Alinhac. Blowup for nonlinear hyperbolic equations, volume 17 of Progr. Nonlinear Differential Equations Appl. Birkhäuser Boston, Inc., Boston, MA, 1995.
- [4] Gianni Arioli, Filippo Gazzola, and Hans Koch. Uniqueness and bifurcation branches for planar steady navier–stokes equations under navier boundary conditions. Journal of Mathematical Fluid Mechanics, 23(3):49, 2021.
- [5] Anxo Biasi. Self-similar solutions to the compressible euler equations and their instabilities. Communications in Nonlinear Science and Numerical Simulation, 103:106014, 2021.
- [6] Tristan Buckmaster, Gonzalo Cao-Labora, and Javier Gómez-Serrano. Smooth imploding solutions for 3d compressible fluids. Forum of Mathematics, Pi, to appear.
- [7] Tristan Buckmaster, Theodore D. Drivas, Steve Shkoller, and Vlad Vicol. Simultaneous development of shocks and cusps for 2D euler with azimuthal symmetry from smooth data. Annals of PDE, 8(2):1–199, 2022.
- [8] Tristan Buckmaster, Steve Shkoller, and Vlad Vicol. Formation of shocks for 2d isentropic compressible euler. Communications on Pure and Applied Mathematics, 75(9):2069–2120, 2022.
- [9] Tristan Buckmaster, Steve Shkoller, and Vlad Vicol. Formation of point shocks for 3d compressible euler. Comm. Pure Appl. Math., 76(9):2073–2191, 2023.
- [10] Tristan Buckmaster, Steve Shkoller, and Vlad Vicol. Shock formation and vorticity creation for 3d euler. Comm. Pure Appl. Math., 76(9):1965–2072, 2023.
- [11] Gonzalo Cao-Labora, Javier Gómez-Serrano, Jia Shi, and Gigliola Staffilani. Non-radial implosion for the defocusing nonlinear Schrödinger equation in and . arXiv preprint arXiv:2410.04532, 2024.
- [12] Gonzalo Cao-Labora, Javier Gómez-Serrano, Jia Shi, and Gigliola Staffilani. Non-radial implosion for compressible Euler and Navier-Stokes in and . arXiv preprint arXiv:2310.05325, 2023.
- [13] Angel Castro, Diego Córdoba, and Javier Gómez-Serrano. Global smooth solutions for the inviscid SQG equation. Mem. Amer. Math. Soc., 266(1292):v+89, 2020.
- [14] G.-Q. Chen and M. Schrecker. Global entropy solutions and newtonian limit for the relativistic euler equations. Ann. PDE, 8:Paper No. 10, 53 pp., 2022.
- [15] Jiajie Chen. Vorticity blowup in compressible Euler equations in . arXiv preprint arXiv:2408.04319, 2024.
- [16] Jiajie Chen, Giorgio Cialdea, Steve Shkoller, and Vlad Vicol. Vorticity blowup in 2d compressible euler equations. arXiv preprint arXiv:2407.06455, 2024.
- [17] Jiajie Chen and Thomas Y Hou. Stable nearly self-similar blowup of the 2D Boussinesq and 3D Euler equations with smooth data II: Rigorous numerics. arXiv preprint: arXiv:2305.05660.
- [18] Jiajie Chen and Thomas Y Hou. Stable nearly self-similar blowup of the 2D Boussinesq and 3D Euler equations with smooth data I: Analysis. arXiv preprint: arXiv:2210.07191v3, 2022.
- [19] Jiajie Chen, Thomas Y Hou, and De Huang. On the finite time blowup of the De Gregorio model for the 3D Euler equations. Communications on Pure and Applied Mathematics, 74(6):1282–1350, 2021.
- [20] Jiajie Chen, Thomas Y Hou, and De Huang. Asymptotically self-similar blowup of the Hou–Luo model for the 3D Euler equations. Annals of PDE, 8(2):24, 2022.
- [21] Demetrios Christodoulou. The formation of shocks in 3-dimensional fluids, volume 2. European Mathematical Society, 2007.
- [22] Demetrios Christodoulou and Shuang Miao. Compressible flow and Euler’s equations, volume 9. International Press Somerville, MA, 2014.
- [23] Marcelo M Disconzi. Recent developments in mathematical aspects of relativistic fluids. arXiv preprint arXiv:2308.09844, 2023.
- [24] R. Donninger. Strichartz estimates in similarity coordinates and stable blowup for the critical wave equation. Duke Math. J., 166:1627–1683, 2017.
- [25] T. Duyckaerts, C. E. Kenig, and F. Merle. Profiles of bounded radial solutions of the focusing, energy-critical wave equation. Geom. Funct. Anal., 22:639–698, 2012.
- [26] T. Duyckaerts, C. E. Kenig, and F. Merle. Classification of radial solutions of the focusing, energy-critical wave equation. Camb. J. Math., 1:75–144, 2013.
- [27] T. Duyckaerts and J. Yang. Blow-up of a critical sobolev norm for energy-subcritical and energy-supercritical wave equations. Anal. PDE, 11:983–1028, 2018.
- [28] Jean Ginibre and Giorgio Velo. The global cauchy problem for the non linear klein-gordon equation. Mathematische Zeitschrift, 189:487–505, 1985.
- [29] Jean Ginibre and Giorgio Velo. The global cauchy problem for the non linear klein-gordon equation-ii. In Annales de l’IHP Analyse non linéaire, volume 6, pages 15–35, 1989.
- [30] Javier Gómez-Serrano. Computer-assisted proofs in pde: a survey. SeMA Journal, 76(3):459–484, 2019.
- [31] M. Grillakis. Regularity for the wave equation with a critical nonlinearity. Comm. Pure Appl. Math., 45:749–774, 1992.
- [32] Manoussos G Grillakis. Regularity and asymptotic behavior of the wave equation with a critical nonlinearity. Annals of mathematics, 132(3):485–509, 1990.
- [33] KG Guderley. Starke kugelige und zylindrische verdichtungsstosse in der nahe des kugelmitterpunktes bnw. der zylinderachse. Luftfahrtforschung, 19:302, 1942.
- [34] Yan Guo, Mahir Hadžić, Juhi Jang, and Matthew Schrecker. Gravitational collapse for polytropic gaseous stars: self-similar solutions. Archive for Rational Mechanics and Analysis, 246(2):957–1066, 2022.
- [35] Yan Guo and A Shadi Tahvildar-Zadeh. Formation of singularities in relativistic fluid dynamics and in spherically symmetric plasma dynamics. Contemp. Math, 238:151–161, 1999.
- [36] Juhi Jang, Jiaqi Liu, and Matthew Schrecker. On self-similar converging shock waves. arXiv e-prints, page arXiv:2310.18483, October 2023.
- [37] J. Jendrej. Construction of type II blow-up solutions for the energy-critical wave equation in dimension 5. J. Funct. Anal., 272:866–917, 2017.
- [38] F. John. Blow-up of solutions of nonlinear wave equations in three space dimensions. Manuscripta Math., 28:235–268, 1979.
- [39] Konrad Jörgens. Das anfangswertproblem in großen für eine klasse nichtlinearer wellengleichungen. Mathematische Zeitschrift, 77(1):295–308, 1961.
- [40] C. E. Kenig. Lectures on the energy critical nonlinear wave equation, volume 122 of CBMS Reg. Conf. Ser. Math. Published for the Conference Board of the Mathematical Sciences, Washington, DC, 2015.
- [41] C. E. Kenig and F. Merle. Global well-posedness, scattering and blow-up for the energy-critical focusing non-linear wave equation. Acta Math., 201:147–212, 2008.
- [42] J. Krieger and W. Schlag. Full range of blow up exponents for the quintic wave equation in three dimensions. J. Math. Pures Appl. (9), 101:873–900, 2014.
- [43] J. Krieger, W. Schlag, and D. Tataru. Slow blow-up solutions for the critical focusing semilinear wave equation. Duke Math. J., 147:1–53, 2009.
- [44] Peter D. Lax. Development of singularities of solutions of nonlinear hyperbolic partial differential equations. J. Mathematical Phys., 5:611–613, 1964.
- [45] H. A. Levine. Instability and nonexistence of global solutions to nonlinear wave equations of the form . Trans. Amer. Math. Soc., 192:1–21, 1974.
- [46] Hans Lindblad. Blow-up for solutions of with small initial data. Communications in partial differential equations, 15(6):757–821, 1990.
- [47] Jonathan Luk and Jared Speck. Shock formation in solutions to the 2d compressible euler equations in the presence of non-zero vorticity. Inventiones mathematicae, 214(1):1–169, 2018.
- [48] T. Makino and S. Ukai. Local smooth solutions of the relativistic euler equation. J. Math. Kyoto Univ., 35:105–114, 1995.
- [49] F Merle and P Raphael. Blow up dynamic and upper bound on the blow up rate for critical nonlinear Schrödinger equation. Journées équations aux dérivées partielles, pages 1–5, 2002.
- [50] Frank Merle, Pierre Raphaël, Igor Rodnianski, and Jeremie Szeftel. On blow up for the energy super critical defocusing nonlinear Schrödinger equations. Inventiones mathematicae, 227(1):247–413, 2022.
- [51] Frank Merle, Pierre Raphaël, Igor Rodnianski, and Jeremie Szeftel. On the implosion of a compressible fluid I: smooth self-similar inviscid profile. Ann. of Math. (2), 196(2):567–778, 2022.
- [52] Frank Merle, Pierre Raphaël, Igor Rodnianski, and Jeremie Szeftel. On the implosion of a compressible fluid II: singularity formation. Ann. of Math. (2), 196(2):779–889, 2022.
- [53] Ronghua Pan and Joel A Smoller. Blowup of smooth solutions for relativistic euler equations. Communications in mathematical physics, 262:729–755, 2006.
- [54] L. Ruan and C. Zhu. Existence of global smooth solution to the relativistic euler equations. Nonlinear Anal., 60:993–1001, 2005.
- [55] Feng Shao, Dongyi Wei, and Zhifei Zhang. On blow-up for the supercritical defocusing nonlinear wave equation. arXiv preprint arXiv:2405.19674, 2024.
- [56] Feng Shao, Dongyi Wei, and Zhifei Zhang. Self-similar imploding solutions of the relativistic euler equations. arXiv preprint arXiv:2403.11471, 2024.
- [57] Jalal Shatah and Michael Struwe. Regularity results for nonlinear wave equations. Annals of Mathematics, 138(3):503–518, 1993.
- [58] Jalal Shatah and Michael Struwe. Well-posedness in the energy space for semilinear wave equations with critical growth. International Mathematics Research Notices, 1994(7):303–309, 1994.
- [59] Steve Shkoller and Vlad Vicol. The geometry of maximal development and shock formation for the Euler equations in multiple space dimensions. Inventiones mathematicae, 237(3):871–1252, Jun 2024.
- [60] J. Smoller and B. Temple. Global solutions of the relativistic euler equations. Comm. Math. Phys., 156:67–99, 1993.
- [61] Christopher Donald Sogge. Lectures on non-linear wave equations, volume 2. International Press Boston, MA, 1995.
- [62] Richard P Stanley. Catalan numbers. Cambridge University Press, 2015.
- [63] Michael Struwe. Globally regular solutions to the klein-gordon equation. Annali della Scuola Normale Superiore di Pisa-Classe di Scienze, 15(3):495–513, 1988.
- [64] T. Tao. Finite-time blowup for a supercritical defocusing nonlinear wave system. Anal. PDE, 9:1999–2030, 2016.
- [65] Terence Tao. Nonlinear Dispersive Equations: Local and Global Analysis, volume 106. American Mathematical Society, 2006.
- [66] Jan Bouwe van den Berg, Maxime Breden, Jean-Philippe Lessard, and Lennaert van Veen. Spontaneous periodic orbits in the navier–stokes flow. Journal of nonlinear science, 31:1–64, 2021.