Loading [MathJax]/jax/element/mml/optable/BasicLatin.js
Research article Special Issues

A characteristics approach to shock formation in 2D Euler with azimuthal symmetry and entropy

  • We provide a detailed analysis of the shock formation process for the non-isentropic 2d Euler equations in azimuthal symmetry. We prove that from an open set of smooth and generic initial data, solutions of the Euler equations form a first singularity or gradient blow-up. This first singularity is termed a Hölder C13 pre-shock, and our analysis provides the first detailed description of this cusp solution. The novelty of this work relative to [1] is that we herein consider a much larger class of initial data, allow for a non-constant initial entropy, allow for a non-trivial sub-dominant Riemann variable, and introduce a host of new identities to avoid apparent derivative loss due to entropy gradients. The method of proof is also new and robust, exploring the transversality of the three different characteristic families to transform space derivatives into time derivatives. Our main result provides a fractional series expansion of the Euler solution about the pre-shock, whose coefficients are computed from the initial data.

    Citation: Isaac Neal, Steve Shkoller, Vlad Vicol. A characteristics approach to shock formation in 2D Euler with azimuthal symmetry and entropy[J]. Communications in Analysis and Mechanics, 2025, 17(1): 188-236. doi: 10.3934/cam.2025009

    Related Papers:

    [1] Anthony Bloch, Marta Farré Puiggalí, David Martín de Diego . Metriplectic Euler-Poincaré equations: smooth and discrete dynamics. Communications in Analysis and Mechanics, 2024, 16(4): 910-927. doi: 10.3934/cam.2024040
    [2] Panyu Deng, Jun Zheng, Guchuan Zhu . Well-posedness and stability for a nonlinear Euler-Bernoulli beam equation. Communications in Analysis and Mechanics, 2024, 16(1): 193-216. doi: 10.3934/cam.2024009
    [3] Shuyue Ma, Jiawei Sun, Huimin Yu . Global existence and stability of temporal periodic solution to non-isentropic compressible Euler equations with a source term. Communications in Analysis and Mechanics, 2023, 15(2): 245-266. doi: 10.3934/cam.2023013
    [4] Hongxia Lin, Sabana, Qing Sun, Ruiqi You, Xiaochuan Guo . The stability and decay of 2D incompressible Boussinesq equation with partial vertical dissipation. Communications in Analysis and Mechanics, 2025, 17(1): 100-127. doi: 10.3934/cam.2025005
    [5] Sixing Tao . Lie symmetry analysis, particular solutions and conservation laws for the dissipative (2 + 1)- dimensional AKNS equation. Communications in Analysis and Mechanics, 2023, 15(3): 494-514. doi: 10.3934/cam.2023024
    [6] Meiqiang Feng . Nontrivial $ p $-convex solutions to singular $ p $-Monge-Ampère problems: Existence, Multiplicity and Nonexistence. Communications in Analysis and Mechanics, 2024, 16(1): 71-93. doi: 10.3934/cam.2024004
    [7] Yao Sun, Pan Wang, Xinru Lu, Bo Chen . A boundary integral equation method for the fluid-solid interaction problem. Communications in Analysis and Mechanics, 2023, 15(4): 716-742. doi: 10.3934/cam.2023035
    [8] Xiulan Wu, Yaxin Zhao, Xiaoxin Yang . On a singular parabolic $ p $-Laplacian equation with logarithmic nonlinearity. Communications in Analysis and Mechanics, 2024, 16(3): 528-553. doi: 10.3934/cam.2024025
    [9] Mustafa Avci . On an anisotropic $ \overset{\rightarrow }{p}(\cdot) $-Laplace equation with variable singular and sublinear nonlinearities. Communications in Analysis and Mechanics, 2024, 16(3): 554-577. doi: 10.3934/cam.2024026
    [10] Chunming Ju, Giovanni Molica Bisci, Binlin Zhang . On sequences of homoclinic solutions for fractional discrete $ p $-Laplacian equations. Communications in Analysis and Mechanics, 2023, 15(4): 586-597. doi: 10.3934/cam.2023029
  • We provide a detailed analysis of the shock formation process for the non-isentropic 2d Euler equations in azimuthal symmetry. We prove that from an open set of smooth and generic initial data, solutions of the Euler equations form a first singularity or gradient blow-up. This first singularity is termed a Hölder C13 pre-shock, and our analysis provides the first detailed description of this cusp solution. The novelty of this work relative to [1] is that we herein consider a much larger class of initial data, allow for a non-constant initial entropy, allow for a non-trivial sub-dominant Riemann variable, and introduce a host of new identities to avoid apparent derivative loss due to entropy gradients. The method of proof is also new and robust, exploring the transversality of the three different characteristic families to transform space derivatives into time derivatives. Our main result provides a fractional series expansion of the Euler solution about the pre-shock, whose coefficients are computed from the initial data.



    Investigating shock formation and development is one of the central problems of hyperbolic PDE. Establishing shock formation (gradient blowup) from smooth initial data, in a constructive manner, is crucial for analyzing the dynamics of the resulting discontinuous shock waves. A precise description of the solution at the pre-shock (the spacetime set where smooth solutions first form cusps) is what allows for a full characterization of singularity propagation, especially in multiple space dimensions (see § 1.2 for details).

    This paper establishes shock formation for smooth solutions of the non-isentropic two-dimensional compressible Euler equations in azimuthal symmetry. When compared to [2] this work gives a detailed description of the solution near the pre-shock as a fractional power series. This paper also goes beyond [1] by establishing shock formation in the non-isentropic setting, and with minimal constraints imposed on the initial data (see § 1.2 for details).

    Beyond the result itself, we develop a new robust proof strategy for establishing shock formation for a complex system of hyperbolic PDEs with multiple wave speeds. Instead of appealing to modulated self-similar analysis (cf. [1,2]), we use new variables that satisfy pointwise and integral identities and accurately capture the compressible Euler dynamics (see § 1.3 for details).

    The Euler equations of gas dynamics consist of the three conservation laws for momentum, mass, and energy, given respectively by

    t(ρu)+div(ρuu+pI)=0, (1.1a)
    tρ+div(ρu)=0, (1.1b)
    tE+div((p+E)u)=0. (1.1c)

    In two space dimensions, the focus of this paper, u:R2×RR2 denotes the velocity vector field, ρ:R2×RR+ denotes the strictly positive density function, E:R2×RR denotes the total energy function, and p:R2×RR denotes the pressure function which is related to (u,ρ,E) by the identity p=(γ1)(E12ρ|u|2), where γ>1 denotes the adiabatic exponent. For the analysis of the shock formation process, it is convenient to replace conservation of energy (1.1c) with transport of entropy

    tS+uS=0. (1.1d)

    Here, S:R2×RR denotes the specific entropy, and the equation-of-state for pressure is written as

    p(ρ,S)=1γργeS. (1.2)

    In preparation for reducing the equations to a more symmetric form, using Riemann-type variables, we introduce the adiabatic exponent

    α=γ12

    so that the (rescaled) sound speed reads

    σ=1αpρ=1αeS2ρα. (1.3)

    With this notation, the ideal gas equation of state (1.2) becomes

    p=α2γρσ2. (1.4)

    The Euler equations (1.1a), (1.1b), and (1.1d), as a system for (u,σ,S), are then given by

    tu+(u)u+ασσ=α2γσ2S, (1.5a)
    tσ+(u)σ+ασdivu=0, (1.5b)
    tS+(u)S=0. (1.5c)

    We let ω=u denote the scalar vorticity, and define the specific vorticity by ζ=ωρ. A straightforward computation shows that ζ is a solution to

    tζ+(u)ζ=αγσρσS. (1.6)

    The term αγσρσS on the right side of (1.6) can also be written as ρ3ρp and is referred to as baroclinic torque.

    The goal of this paper is to give a constructive proof of shock formation for (1.5), from smooth initial data, via a method powerful enough to capture a high-order series expansion of all fields at the preshock, information which is in turn necessary to study the shock development problem. More precisely, we prove:

    Theorem 1.1 (Main result, abbreviated). From smooth, non-isentropic initial data with azimuthal symmetry lying in an open set*, there exist smooth solutions to the 2d Euler equations (1.1) that form a gradient blowup singularity at a computable time T and spatial location. More specifically, there exists ξT such that when the 2d Euler equations are expressed in polar coordinates as in (2.1), the azimuthal component of the flow uθ and the sound speed σ form C0,13 cusps along the ray θ=ξ at the time of the blowup, and are given by the fractional series expansions

    *See § 2.3-2.4 for the details of the pertinent set of initial data.

    We abuse notation here, because the time T used here differs from the time T referenced in the rest of the paper by a constant dependent on γ>1. See § 2.1.

    uθ(r,θ,T)=r(b0+b1(θξ)1/3+b2(θξ)2/3+O(ε1|θξ|)),σ(r,θ,T)=r(c0+b1(θξ)1/3+b2(θξ)2/3+O(ε1|θξ|)),

    for θ in a neighborhood of radius ε3, while the radial component ur of the flow, the specific entropy S, and the specific vorticity ζ remain C1,13, with fractional series expansions

    Here ε1 is a large parameter quantifying the absolute size of slope of the initial data. See § 2.3 for details.

    ur(r,θ,T)=r(å0+å3(θξ)+å4(θξ)4/3+O(ε1/2|θξ|5/3)),S(r,θ,T)=k0+k3(θξ)+k4(θξ)4/3+O(ε1|θξ|5/3),ζ(r,θ,T)=v0+v3(θξ)+O(ε1|θξ|4/3).

    Here, the constants å0,å3,å4,b0,b1,b2,c0,k0,k3,k4, and v0 are O(1) while v3 is O(ε1).§

    §See § 2.2 for the details of our use of O() and .

    We recall that the classical proofs of finite-time singularity formation for the compressible Euler equations and related hyperbolic systems are not constructive (see e.g. [3,4,5] for small smooth perturbations near constant states, and [6,7] for large data). We refer the reader to [8,9,10] for an extensive bibliographic account of this classical theory.

    A constructive proof of blowup, and equally importantly, a detailed description of the solution at the pre-shock, is necessary in order to establish shock development. By definition, shock development refers to the instantaneous development of the discontinuous shock wave from the C0,13 Hölder cusp at the pre-shock. This is especially true in multiple space dimensions: while the theory of weak solutions for 1D hyperbolic systems is well-developed (see e.g. [8]), many of the techniques used in the 1D theory either do not apply in multiple space dimensions or are not precise enough to be useful for the shock development problem, which requires bounds on derivatives of the solution.

    For example, the BV estimates utilized in the classical theory of shocks for 1D hyperbolic systems fails for d2. See [11].

    Departing from the weak solutions perspective, Lebaud [12] established shock formation and development for the one-dimensional p-system (a variant on 1D isentropic Euler). These results were expanded upon by Chen and Dong [13] and Kong [14]. Studying shock development in the p-system does not per se prove anything about physical solutions of Euler, because physical solutions of Euler that have shocks cannot be isentropic (see § 2.2 of [1] or § 3 of [15] for details). Moreover, non-isentropic solutions of Euler are generically not irrotational due to a misalignment of pressure and entropy gradients (see (1.6) above and § 4 of [15] for the 3D case), so physical solutions which have shocks are also generically not irrotational. Studying shock development for piecewise isentropic or even piecewise irrotational solutions of Euler is called the restricted shock development problem. For the restricted shock development problem, Christodoulou established shock formation and development for irrotational flows in his landmark books [16,17]. Yin [18] wrote the first paper establishing shock formation and development for non-isentropic Euler, but confined to spherical symmetry (see also [19]). Luk and Speck [20] proved shock formation for the 2D isentropic Euler equations in the presence of vorticity by an extension of Christodoulou's geometric framework to allow for vorticity transport. In [21], they later generalized their 2D result to the full 3D non-isentropic setting.

    A different perspective was taken by Buckmaster, Shkoller, and Vicol [2,22,23], who used modulated self-similar variables to construct the first gradient singularity (a point shock) from generic smooth initial data. In [2] they constructed shocks for 2D isentropic Euler in azimuthal symmetry and characterized the shock profile as an asymptotically self-similar, stable 1D blowup profile. After that, they proved for the first time that the 3D isentropic Euler equations generically form a stable point shock, even in the presence of vorticity [22]. The important generalization to the full non-isentropic setting was achieved in [23], where it is also shown that irrotational data instantaneously creates vorticity due to baroclinic torque, and the vorticity remains uniformly bounded up to and including the time of the first gradient singularity.

    The analysis of the Euler evolution beyond the time of the first gradient blowup was recently addressed by Shkoller and Vicol [24] by studying the so-called Maximal Globally Hyperbolic Development (MGHD) of smooth and compressive Cauchy data. This can be understood to be the largest (local) spacetime that contains a smooth (and invertible) evolution of the Cauchy data. The future temporal boundary of this spacetime consists of the codimension-2 manifold of pre-shocks (containing the spacetime set of first gradient catastrophes), the singular set (a downstream hypersurface of gradient blowups emanating from the pre-shock manifold), and the Cauchy horizon (an upstream hypersurface emanating from the pre-shock set which the smooth Euler solution can never cross). A partial construction of the MGHD was also obtained by Abbrescia and Speck [25] who were able to evolve the Euler solution up the union of the pre-shock set and the singular set (but the upstream evolution up to the Cauchy horizon was not treated).

    Buckmaster, Drivas, Shkoller, and Vicol [1] established for the first time shock developement in the presence of voriticity, by working in azimuthal symmetry. By improving upon [2], the solution at the pre-shock is described in [1] by a fractional series, assuming that the flow is initially isentropic (k00 in (2.5c) below) and that the subdominant Riemann variable vanishes (z00 in (2.5b) below). They then used this detailed description of the solution to establish shock development for 2D Euler within the class of azimuthal solutions. The paper [1] is the first to also confirm the production of both a discontinuous shock wave and two surfaces of cusp singularities emanating from the pre-shock, as predicted by Landau and Lifschitz [26].

    This paper breaks with [2] and [1] by forgoing the use of self-similar variables. Instead, we use only the fine structure of the Euler system written in the characteristic coordinates that correspond to the three different wave speeds present in the system. We show that the sound speed remains bounded from below up to the time of the first blowup (see Proposition 4.1), which means that the three wave speeds remain uniformly transverse to one another up to the blowup time. This transversality allows us to prove useful integral bounds (see Lemma 3.1 and § 4) and allows us to exchange space derivatives for time derivatives (see § 5), which can be integrated to obtain identities for the higher-order derivatives of our variables. This exchange of space for time derivatives via transversality is the key new idea of this work.

    The implementation of this idea is made possible by using the special differentiated Riemann variables introduced in [1]. These new variables, labeled qw and qz, evolve along the characteristics of the fastest and slowest wave speeds, respectively, and they do not experience derivative loss (see § 3 of [1] or § 3.2 below). Whereas [1] utilized qw and qz for studying shock development, we use qw and qz to also establish shock formation in the non-isentropic setting. Using pointwise and integral identities for qw and qz, we are able to obtain estimates for our variables and their derivatives up to the blowup time without first establishing the uniqueness or location of the blowup label; we instead derive the uniqueness and location of the blowup label as a result of our estimates (see § 10.3).

    We note that because we avoid self-similar analysis, we are able to place far fewer assumptions on our initial data than in [1]. When compared to [1], we also obtain a higher-order fractional series expansion of the solution at the time of blowup (see Theorem 2.1).

    The 2D Euler equations (1.5) take the following form in polar coordinates for the variables (uθ,ur,ρ,S):

    (t+urr+1ruθθ)ur1ru2θ+ασrσ=α2γσ2rS, (2.1a)
    (t+urr+1ruθθ)uθ+1ruruθ+ασrθσ=α2γσ2rθS, (2.1b)
    (t+urr+1ruθθ)σ+ασ(1rur+rur+1rθuθ)=0, (2.1c)
    (t+urr+1ruθθ)S=0. (2.1d)

    We introduce the new variables||

    ||Note that our symmetry constraints make S discontinuous at the origin unless S is constant. For this reason, a classical solution of the 2D Euler equations (1.5) is recovered from the azimuthal variables (a,b,c,k) via (2.2) on the punctured plane. Alternatively, we may restrict the domain of evolution for 2D Euler to an annular domain pushed forward under the flow of u (see [2, § 2.1]).

    uθ(r,θ,t)=rb(θ,t),   ur(r,θ,t)=ra(θ,t),   σ(r,θ,t)=rc(θ,t),   S(r,θ,t)=k(θ,t). (2.2)

    The system (2.1) then takes the form

    (t+bθ)a+a2b2+αc2=0 (2.3a)
    (t+bθ)b+αcθc+2ab=α2γc2θk (2.3b)
    (t+bθ)c+αcθb+γac=0 (2.3c)
    (t+bθ)k=0. (2.3d)

    For simplicity of the presentation, we will set γ=2 from here on; note however that all statements in this paper apply mutatis mutandis to the case of a general γ>1. The Riemann functions w and z are defined by

    w=b+c,z=bc, (2.4a)
    w=b+c,z=bc,b=12(w+z),c=12(wz). (2.4b)

    It is convenient to rescale time, letting t34˜t, and for notational simplicity, we continue to write t for ˜t. With this temporal rescaling employed, the system (2.3c) can be equivalently written as

    tw+λ3θw=83aw+124(wz)2θk, (2.5a)
    tz+λ1θz=83az+124(wz)2θk, (2.5b)
    tk+λ2θk=0, (2.5c)
    ta+λ2θa=43a2+13(w+z)216(wz)2. (2.5d)

    where the three wave speeds are given by

    λ1=13w+z<λ2=23w+23z<λ3=w+13z. (2.6)

    We note that (2.3c) takes the form

    tc+λ2θc+12cθλ2=83ac. (2.7)

    Finally, we denote the specific vorticity (1.6) in azimuthal symmetry by

    ϖ=4(w+zθa)c2ek, (2.8)

    which satisfies the evolution equation

    tϖ+λ2θϖ=83aϖ+43ekθk. (2.9)

    In most of what follows, there will be an important parameter ε>0, and ab will be used to signify that aCb for some constant C independent of ε and any variables x,θ, or t. However, the constant can depend on the implicit constants in the assumptions on the initial data in § 2.3 and can depend on our choice of γ>1 for the pressure law**. We will use the notation ab to express aba. We will also write

    **We have already chosen to fix γ=2 for the entirety of this paper, but our result will hold for arbitrary γ>1, and the value of γ will effect the constants.

    f=O(g)

    to express that |f|g everywhere in the relevant domain. We will express bounds of the type

    f(x,t)={O(b1)|x|ε2O(b2)|x|ε2simply asf=B(b1;b2).

    Often below we will have functions f defined on T×[0,T) and maps Ψ:T×[0,T)T, and we will use the notation

    fΨ(x,t):=f(Ψ(x,t),t).

    When such an inverse exists, we will write Ψ1 to denote the function such that Ψ1Ψ(x,t)=ΨΨ1(x,t)=x for all t.

    While the spatial variable θ for (2.5) lies in T, we will often identify T with the interval (π,π].

    Our initial data will be w0,z0,k0,a0H6(T), where z0,k0, and a0 all satisfy

    jxk0Lεγj,jxa0Lεαj,jxz0Lεβj, (2.10)

    for j=0,1,2,3,4,5, where αj,βj,γj are fixed constants satisfying the relations

    α0,β0,γ00,

    γ1μ, α10,

    γjμj for j=2,3,4,5,

    αjμ+1j for j=2,3,4,5,

    βjμj for j=1,2,3,4,5.

    Here μ>0 is a fixed positive constant that is a lower bound on the distance of our vector of parameters (α2,,α5,γ1,,γ5,β1,,β5) from the boundary of the open set defined by the constraints β1>1, γ1>0, etc. Additionally, we assume that w0 satisfies

    1. w01,

    2. w0(0):=1ε and |w0(x)|<ε1 for all x0,

    3. w0(x)1ε+Cεμ21 for all |x|ε3/2, and some constant C>0.

    4. w0(x)ε4 for all |x|ε3/2,

    5. |4xw0(x)|εμ5 for all |x|ε2,

    6. 5xw0Lε7,

    and that z0 satisfies

    max (2.11)

    Note that an immediate consequence of our assumptions is that must also satisfy

    ,

    for ,

    ,

    ,

    .

    The following additional constraints are not at all necessary for proving our theorem, but they do make the formulas of the proof below cleaner

    (2.12)

    Note that the constraints made here on the first five derivatives of are much less stringent than those imposed in [1]. In [1], the authors assume that is constant, is identically 0, and that and have support with diameter , among other constraints. Here we do away with such unnecessary hypotheses. Additionally, the result of this paper applies to a wide range of parameters , whereas in [1] the authors only work with , which is only one point in our admissible range for these parameters.

    In what follows, we will parametrize time so that the initial time is always . The local well-posedness theory of (1.5) implies that for any there exists a time such that there exists a unique solution of (2.5) satisfying . Furthermore, is guaranteed to be in . Additionally, it follows from the standard theory of (1.5) that if then

    (2.13)

    The inequalities above can be made into open constraints by making them strict inequalities. While the two pointwise constraints that require to attain its unique global minimum at and are not open constraints, for any suitably small perturbation of initial data satisfying all of the above constraints, one can recover the two pointwise constraints by translating in space and rescaling the solution in time. Since the spatial translation and time rescaling can be made sufficiently small, there exists an open set of initial data around the functions described above for which the results of Theorem 2.1 below still hold. Thus, the shock formation we describe is stable.

    Theorem 2.1 (Main theorem). For , sufficiently small, and initial data in the open set described in § 2.3, there exists a blowup time with , a unique blowup location , and unique solutions to (2.5) on such that ,

    where is the specific vorticity (see (2.8)). Furthermore, there exists a unique blowup label such that

    where is the 3-characteristic defined in § 3.1. In a neighborhood of radius the functions , and have the following fractional series expansions:

    There exist constants with

    such that

    (2.14)

    There exist constants with

    such that

    (2.15)

    There exist constants with

    such that

    (2.16)

    There exist constants with

    such that

    (2.17)

    There exist constants with

    such that

    (2.18)

    Moreover, the regularity away from the pre-shock is characterized by

    (2.19)

    Theorem 1.1 clearly follows from Theorem 2.1 as an immediate corollary.

    In this paper, we will show that the classical solution of (2.5) with the initial data specified in § 2.3 breaks down in finite time, and that this occurs when the flow of the fastest wave speed ceases to be a diffeomorphism. More specifically, the blowup time will be characterized as the first time when . We will also establish that there is a unique Lagrangian label for which , which will imply that vanishes at as well. While and will be shown to remain bounded on , will be shown to go to at the point and remain smooth elsewhere. The key ingredient for implementing the above-described strategy is to show that the functions , and remain as smooth as their initial data, uniformly up to . The authors of [1] proved such uniform estimates using self-similar analysis, but only in a special case.†† In this paper, we prove uniform estimates for on , even in the most general setting, not by relying on self-similar variables, but by instead using the transversality of various families of characteristics. This allows us to also consider a much broader class of initial data than previously considered in [1]. Once we have shown that all the variables stay smooth along the characteristic, we obtain our functional description of the solution near by inverting the map for near the point . In light of the constraints , this amounts to the inversion of what is to leading order a cubic polynomial, resulting in fractional series expansions of and near in terms of powers of .

    ††The authors of [1] work in the case where and are identically zero and many more constraints are placed on and . See § 2.3 above for a discussion.

    This paper is organized as follows:

    1. In § 4 we bound and prove that must become infinite at time . We use a simple bootstrap argument to get estimates for and their first derivatives up to time . Using these estimates, we show that must have a zero before time , and conclude that . This implies that and therefore all of our estimates and identities hold up to time . The fact that must blow up then follows immediately from the fact that and remain bounded up to time (see (2.13) above).

    2. Next we show that and remain smooth up to time . To do this, we first establish crucial identities in § 5, which result from the fact that the wave speeds are uniformly transverse to one another. Then in § 6 - 9 we prove pointwise bounds on and their derivatives in terms of and its derivatives by analyzing how our new variables evolve along the multiple wave speeds. This allows us to conclude in § 10 that and all remain smooth along .

    3. Last we establish that the singularity occurs at a unique point and we invert near this point to obtain fractional series expansions for and . We do this by establishing in § 10 that there is a unique point where vanishes and that as well. Since near , it follows (see § 11) that for small at time , and the Taylor series expansions of the smooth functions , and near become fractional series expansions of and near .

    Let , and let be the flow of .

    (3.1)

    If everywhere, this tells us that

    (3.2)

    If and are bounded, then this lets us conclude that . We will prove in the next section that and that are indeed bounded on , so everything that follows is relevant.

    In the case where , we have , the first wave speed. Let denote the corresponding flow, the so-called 1-characteristic. Its first derivative satisfies

    (3.3)

    while its second derivative obeys

    (3.4)
    (3.5)

    When , we have and the corresponding flow is the 2-characteristic, . The first derivative of satisfies

    (3.6)

    while its second derivative obeys

    (3.7)
    (3.8)

    When , we have and the corresponding flow is the 3-characteristics, . Note that our analysis for breaks down for , but also that is essentially transported along .

    Our system (2.5) can be written as

    (3.9)

    where

    Taking of (3.9) and diagonalizing gives us

    where , , and is a third-order polynomial. This motivates the introduction of the following variables:

    (3.10)

    On can check using the identities in § A.1 that

    (3.11)
    (3.12)

    If we define

    (3.13)

    then our equation for gives us the Duhamel formula

    (3.14)

    It follows immediately from the definitions of and that

    (3.15)

    Identity (3.16) will be used in § 4.3, and (3.14) and (3.15) will be used in § 3.3, 4.3, 4.4, and 10. Similarly, satisfies the Duhamel formula

    (3.16)

    Let , and let be the flow of .

    Lemma 3.1. Suppose that and that for all we have

    Then, for all , we have

    (3.17)

    with a constant uniform in

    Proof of Lemma 3.1. Fix and define

    We compute that

    (3.18)

    Note that everywhere. We also know that

    Our hypotheses allow us to conclude (see (4.2) below) that

    for times . Using our hypotheses, along with this equation and (3.14), we conclude that for all , we have

    Plugging this into (3.15) and using the fact that gives us

    The last inequality is true for taken to be small enough, since . Plugging this into (3.14) and letting be sufficiently small gives us

    (3.19)

    It follows that

    Since , it follows that

    So

    Our result follows immediately from this inequality and our hypotheses.

    Proposition 4.1. For small enough, the following estimates hold for all :

    Proof of Proposition 4.1. This follows from an easy bootstrap argument. Let . If we assume all of the listed bounds hold up to time for some constants, then it follows that (3.6) holds up to time . satisfies and

    (4.1)

    Additionally, (2.5) gives us Duhamel formulas for , and . Using these Duhamel formulas along with (2.11), (3.6), (4.1), and the fact that it is straightforward to improve our bounds for all times before , provided the constants we assumed in our bootstrap hypothesis are appropriate and is small enough.

    Using these estimates, it is easy to show that for sufficiently small we obtain that

    (4.2)

    Using (3.6) and (4.1) we have

    (4.3)

    where

    (4.4)

    Therefore,

    (4.5)

    Note that

    (4.6)

    This relation will be useful for estimating the higher derivatives of .

    Since

    (4.7)

    our assumptions on our initial data let us conclude that , and therefore for all we have

    (4.8)

    Since

    (4.9)

    it follows that

    (4.10)

    for all times .

    Using (4.10) and the bounds on the initial data, we conclude that

    (4.11)

    for all times .

    Proposition 4.2. For all we have

    Proof of 4.2. We will use a bootstrap argument. Let and let our bootstrap assumption be that

    for all and a constant to be determined. Since we are assuming for times , it follows from (3.3) and our estimates from § 4.1 that with constants independent of for times , provided that is small enough relative to .

    Using our bootstrap assumption, along with the estimates from § 4.1 and 4.2, we can conclude (see Lemma 3.1 and its proof) that for all , we have

    Using this last estimate along with the estimates from § 4.1, 4.2, it follows from (3.16) and the fact that that

    for times . It follows that

    for . Since , it follows that if we let become small enough, we obtain that

    for all . If is chosen large enough and is chosen small enough, this improves upon our second bootstrap assumption.

    It follows as an immediate corollary of this proposition that

    (4.12)
    (4.13)
    (4.14)

    for all times .

    Now our estimates will let us conclude that behaves roughly the same as it would if were the solution of Burger's equation with initial data and were the flow of . Using Proposition 4.1, (4.2), (4.10), (4.12), and (4.13) in equation (3.14) gives us

    for all times . Plugging this into (3.15) and using the same bounds produces

    (4.15)

    for . Evaluating (4.15) at and using (4.2) gives

    Since this is true for all , it follows that we must have if is chosen small enough. Therefore, , and everything we have proven for is true for .

    We can also prove a lower bound on . Since for all , it follows from (4.15) and (4.2) that

    everywhere. Therefore, , else and would all stay bounded up to .

    We can also obtain a lower bound for away from 0. Indeed, since for , we have

    (4.16)

    Using Lemma 3.1 and the estimates proven in this section, we can now conclude that the bound (3.17) holds for all This fact will be used so frequently in the rest of the paper that we will not bother to cite it.

    Let and let be the flow of .

    Therefore, if is any differentiable function, we have

    This gives us the following equation:

    (5.1)

    The last term in this expression motivates the following definition:

    Definition 5.1 (Transversality). A differentiable function is transversal (or 1-transversal) if it is bounded and there exist both a constant and bounded functions and such that

    Here , as in the above discussion. If in addition and are themselves transversal functions, we say that is 2-transversal. We recursively define to be -transversal if and are -transversal.

    A few remarks about transversal functions:

    ● If satisfies the transversality condition for one , then it satisfies the transversality condition for all . If indeed, if we have

    for some then for any other we have

    Since is bounded, still satisfies the transversality condition for , albeit with a different choice of bounded function . So the notion of a transversal function is independent of our choice of .

    ● Note that while being transversal does not depend on the choice of (as the previous bullet illustrated), and and are independent of , the function changes based on .

    ● If is a bounded function with bounded derivatives, then is trivially transversal, with and .

    ● If functions are -transversal, then is -transversal. Indeed, we have

    ● If functions are -transversal, then their product is -transversal. Indeed, we have

    ● If is -transversal and then is also -transversal. Indeed,

    ● If is smooth and is -transversal, then is -transversal. Indeed, we have

    This rule will be especially useful for .

    is transversal with and when . It follows inductively that if is -transversal, then is -transversal. At this point, we already know that is at least 1-transversal because it is uniformly , so is currently proven to be at least 2-transversal. , so is also 2-transversal. The fact that both and are transversal was the main ingredient used in the computation of (5.1).

    The following lemma will be used in § 7.3, § 8.3, and § 9.3.

    Lemma 5.2 (Identities for transversal functions along 1-characteristics). If is transversal with

    then we have

    (5.2)

    and

    (5.3)

    From these two equations. we obtain the bounds

    (5.4)

    and

    (5.5)

    Proof of Lemma 5.2. (5.2) follows immediately from (5.1). To prove (5.3),

    The inequalities follow immediately from the equations and the first-order estimates.

    The following lemma will be used in § 7.2, § 8.2, and § 9.2.

    Lemma 5.3 (Identities for transversal functions along 2-characteristics). If is a differentiable function satisfying the transversality condition

    then we have

    (5.6)

    and

    (5.7)

    Proof of Lemma 5.3. (5.6) follows immediately from (5.1). The proof of (5.7) is an easy computation using (5.1) and (3.7).

    The following lemma will first be used in § 8.3, so there is no circularity in its proof. See § 6.1 for the definition of and § 6.3 for the definition of .

    Lemma 5.4 (Identities for 2-transversal functions along 1-characteristics). If is 2-transversal with

    then we have

    (5.8)

    and

    (5.9)

    Proof of Lemma 5.4. Taking of (5.2) gives us

    If we define , then the rules for transversal functions tell us that

    Applying (5.3) to and simplifying gives us (5.8).

    For the next identity, we see that

    Applying (5.3) to the function and simplifying gives us (5.9).

    The following lemma will be used in § 8 and § 9.

    Lemma 5.5 (Classes of transversal functions). Let , and let be the flow of . Then

    1. If is a transversal function and is defined by

    then is transversal.

    2. is a transversal function.

    3. If is a transversal function and is defined by

    then is transversal.

    4. If is a 2-transversal function and is defined by

    then is 2-transversal.

    5. is a 2-transversal function.

    6. If is a 2-transversal function and is defined by

    then is 2-transversal.

    Proof of Lemma 5.5. In this proof, satisfies

    (i) Since

    (5.10)

    it follows from (3.17) and the fact that that is transversal.

    (ii) We know that is transversal, and it will be proven in § 6.3 that is transversal. Therefore, part (ⅰ) applies to the function

    Since is smooth, it follows that

    is transversal. We already know that and are both 2-transversal, so it now follows from (3.2) that is transversal.

    (iii) Using (5.1) tells us that

    (5.11)

    It now follows from (6.18) that is transversal.

    (iv) This follows immediately from applying (ii) and (iii) to (5.10).

    (v) It will be proven in § 6.2 that is transversal, from which it will follow that is 2-transversal, and it will be proven in § 7.3 that is 2-transversal. Since is smooth, it follows from (iv) that

    is 2-transversal. Since is 2-transversal, it now follows from (3.2) that is 2-transversal.

    (vi) We prove in § 6.2, 6.3, 7.1, and 7.3 that and are all transversal, so our result follows from applying (i), (ii), and (iii) to (5.11).

    Differentiating (4.1) and plugging in (3.7) gives us

    (6.1)

    If we define

    (6.2)

    then it follows that and from (4.11) we conclude that

    (6.3)
    (6.4)

    With this notation, we can write

    (6.5)

    so is transversal. The fact that is transversal will be used through § 7 and § 6.

    Using (3.6) and (A.1a) we have

    (6.6)

    Therefore, differentiating (4.5) gives us

    (6.7)

    It is easy to check that

    (6.8)

    It follows from this bound and (6.7) that

    By differentiating the equation (4.7) and using our assumptions on the initial data, we conclude that . Therefore,

    (6.9)

    Differentiating (4.9) in space and using our first derivative estimates along with (6.9) gives us

    (6.10)

    So (6.9) lets us conclude that is transversal, which will be used in § 7 and § 8. This equation for and our estimate (6.9) also lets us conclude that

    (6.11)

    It now follows from (6.11) that

    (6.12)

    Let us introduce the new variable

    (6.13)

    Using the identities from § A.1 along with (6.2) gives us

    (6.14)

    If we define

    (6.15)

    then (6.14) gives us the Duhamel formula

    (6.16)

    It now follows from (6.4), (6.3), and (6.11) that

    (6.17)

    It follows immediately from this bound and (6.4) that

    (6.18)

    This bound tells us that

    (6.19)

    We can also conclude that is transversal. Indeed

    (6.20)

    The fact that is transversal will be used in § 7 and § 8.

    Using the fact that , we can compute that

    (7.1)

    Define

    (7.2)

    We know from (4.11) and (6.12) that

    (7.3)

    Since

    (7.4)

    it follows that is transversal and therefore is 2-transversal.

    Taking of (6.2) gives us

    (7.5)

    It follows that

    (7.6)

    Taking of (6.7) gives us

    It is easy to use (3.7) and (6.11) to obtain that Using (6.6), we have that

    (7.7)

    Using (6.6) and (7.7), we have that

    Therefore, we have

    (7.8)

    Rearranging this and using , (4.1), (4.6), and (3.8), we have

    (7.9)

    where

    (7.10)
    (7.11)

    Taking two derivatives of (4.7), we find that

    (7.12)

    It follows that

    (7.13)

    We therefore conclude that

    (7.14)

    Differentiating (4.3) and using (4.1) allows us to compute that

    (7.15)

    where

    (7.16)

    This, along with (7.14), implies that is transversal. Therefore (see (6.10)), is 2-transversal.

    Now taking or (6.10) and using (6.4), (6.18), and (7.14) gives us

    (7.17)

    Now, one can compute that

    (7.18)

    It now follows from (4.11), (5.7), (6.4), (6.18), and (7.14) that

    (7.19)

    Therefore,

    (7.20)

    These equations will be used in § 8.1, § 8.2, § 9.1, and § 9.2.‡‡

    ‡‡For the fifth-order estimates, one actually has to write out the full formula for (7.19) and (7.20) and work with it. We will omit such straightforward but space-consuming details.

    We know that

    Recall from our Duhamel formula for that

    (7.21)

    Taking of (7.21) and using (5.4) and (7.3), we have that

    (7.22)

    Therefore

    (7.23)

    Taking of (6.20) and using these bounds, we conclude

    (7.24)

    Since (6.14) can be rewritten as

    we can also conclude that is transversal, and therefore is 2-transversal.

    Taking of and using (4.1) and (7.20) gives us

    Define

    (8.1)

    Then (4.11) and (6.12) tell us that

    Using (7.18), we compute that

    (8.2)

    so is transversal, and therefore is 3-transversal. This will be used in § 9.

    Taking of (7.5) gives us

    (8.3)

    Note that the terms of order happen to cancel when this computation is done.

    It follows from (6.3), (6.18), (7.3), and (7.24) that

    (8.4)

    At this point, we can apply Lemma 5.5 to conclude that the variable defined in (3.8) is transversal. In fact, Lemma 5.5 allows us to conclude that is 2-transversal, but we will not need to use that until § 9.

    Recall from § 7.2 that

    where are defined by (7.10), (7.11), and (7.16). Since are all transversal, it is immediate that and are transversal. We now also know that and are transversal, so Lemma 5.5 lets us conclude that is transversal. So is 2-transversal. This fact will be utilized in § 8.3.

    It is immediately clear from (7.10) that

    Since

    we conclude that

    (8.5)

    So, to estimate , all that remains is to bound .

    We know from (7.19) that

    (8.6)

    We know from (5.6) that

    It is straightforward to compute that

    Taking of (4.7) produces

    (8.7)

    Therefore, taking of (7.11) and using (7.3) and (7.23), we conclude that

    (8.8)

    Therefore,

    (8.9)

    Now, taking of (4.9) and using (6.4), (6.9), (7.6), (7.14), (7.24), and (8.9) shows that

    (8.10)

    Abusing notation, introduce a function defined so that

    Lemma 5.5 implies that is 2-transversal. Using the new function , we can rewrite (7.21) as

    Given everything that has been proven up to this point, and are both 2-transversal. It follows from Lemma 5.5 that is 2-transversal, and since and are both 2-transversal, it thus follows that is 2-transversal. This will be utilized in § 9.

    Using Lemma 5.4 on the functions and , along with estimates from the previous sections, we conclude that

    It therefore follows that

    We conclude that

    It now follows that

    (8.11)

    We already know (see § 8.1) that is 3-transversal, and we will not need to show that is 4-transversal. is 4-transversal, but it doesn't matter for our purposes. One can easily establish the bound

    It now follows from taking of (8.3) that

    (9.1)

    Since and are all 2-transversal, it follows immediately that and are 2-transversal. and are also 2-transversal, so it follows from Lemma 5.5 that is 2-transversal. Therefore, is 3-transversal. This will be used in § 9.3.

    Taking of (8.5) and using the bounds we already have on , and gives us

    can be bounded in a manner similar to the way was bounded. One simply needs to use a lemma similar to Lemma 5.4 but for the 2-characteristics, which is very straightforward to prove at this point. Then, since and can be explicitly computed and bounded §§, one can bound and conclude that

    §§One must write out the full equation for (7.19) in order to do this, which is arduous but straightforward.

    From here, taking of (6.10) gives

    (9.2)

    One can use Lemma 5.2, Lemma 5.4, and Lemma 5.5 to derive a lemma for 3-transversal functions analogous to Lemma 5.4. Bounding now follows in a manner completely analogous to § 8.3. One obtains

    (9.3)

    It follows from the first derivative estimates that

    where is the integrating factor in (3.13). It follows from the second derivative estimates that

    Taking of (3.14) and using these bounds, we find that

    Taking of (3.15) and plugging in this estimate gives us

    (10.1)

    It follows that

    (10.2)

    Plugging (10.1) into our bound for gives us

    (10.3)

    Using the bound along with our second derivative bounds, we obtain that

    (10.4)
    (10.5)
    (10.6)
    (10.7)

    Since

    it now follows that

    Last since

    we know that

    and therefore (10.1) and (10.4) imply that

    (10.8)

    Since

    it follows that

    (10.9)
    (10.10)

    Using the third derivative estimates and (10.9) gives us

    These estimates will be useful in § 10.4 and § 10.5.

    Multiplying the above bounds by gives us

    Using this, we compute that

    This is true for all .

    Taking of (3.15) and using this bound tells us that

    everywhere. Using this bound, we conclude that

    Since for and , this bound lets us conclude that

    and

    We now conclude that

    We know that for all we have

    Therefore, we have the bounds

    Since

    It follows that

    (10.11)

    Last it is easy to use the bounds on and to conclude that

    Lemma 10.1 (Existence and uniqueness of blowup label). There exists a unique label such that . Furthermore, we have and

    Proof of Lemma 10.1. Due to (4.16), we know that is bound below outside of . We know that in , so for all there is at most one zero of in .

    We know from § 10.1 that for all we have

    Recall that for . It follows that for we have

    Since and for , it follows that

    for . Therefore, we have

    It follows that there exists a constant such that for we have and for we have . So for all , there exists a unique zero of in .

    Therefore, we conclude that there exists a curve such that

    Furthermore, we know that for all . From here it is easy to conclude that for and for , so that must be the minimizer of over .

    Define . We know that as and is bound below for , so as . Our result now follows.

    We can now improve upon our lower bounds for . Let be the curve from the proof of Lemma 10.1. If and , there exists in between and such that

    Since , if , then and , so we have

    It follows that for , we have . We already know (see § 4.4) that

    for all , so we conclude that

    (10.12)

    Lemma 10.2 (Improved estimates for and ). There exist constants such that for all , we have

    (10.13)
    (10.14)
    (10.15)

    Proof of Lemma 10.2. Fix a point . We know that is on and is on . Therefore, Taylor's theorem tells us that there exists a point on the segment connecting to such that

    (10.16)

    Similarly, there exists a point on the segment such that

    (10.17)

    We know that

    We also know that since , , and we have

    We also know that for

    So for

    Also

    Our result now follows.

    Using Lemma 10.2, we can now conclude that

    (10.18)
    (10.19)

    The bound (10.18) will let us deduce (2.19), and (10.19) will be used frequently in § 10.4 and § 10.5.

    We know that

    (10.20)

    Therefore,

    It now follows that

    (10.21)
    (10.22)
    (10.23)
    (10.24)

    The usual argument for bounding derivatives of and now gives

    and

    In the end, we obtain that

    (10.25)
    (10.26)
    (10.27)
    (10.28)
    (10.29)

    These estimates are different from the previous sections because they require more algebra and hinge on admittedly unexpected cancellation. First, note that

    Next, note that

    Combining these identities and our estimates gives us

    Therefore,

    The exact same cancellation occurs for the other two variables to give us

    Similar computations prove that

    Now the usual method for bounding the derivatives of and produces

    (10.30)

    In the end, we obtain that

    Using similar computations to those in this section, one can compute that

    This bound, together with similar bounds that we proved for and , combines with (10.18) to establish (2.19).

    In this section, we will confine our attention to labels with .

    Since is , it has the following Taylor expansion about :

    (11.1)

    Here

    (11.2)

    The flow also has the Taylor expansion

    (11.3)

    where , , ,

    (11.4)

    Here , , and . Note that .

    Let . Lemma A.3 implies that there exists a constant such that for all such that we have

    (11.5)
    (11.6)

    A quick bootstrap argument lets us conclude that this formula holds for all . Furthermore, it is easy to show that there exist two constants such that

    So we are working in a neighborhood of radius around .

    If we define

    then we have

    and

    (11.7)

    Squaring (11.5) and cubing (11.6) gives us

    Therefore,

    (11.8)

    Using this formula, one can compute that

    Since and , it follows that

    (11.9)

    Since , it follows that at time we have

    This is the expansion for in Theorem 2.1.

    Now consider

    Differentiating (11.3) twice and using our above expansions for and gives us

    (11.10)

    Using the fact that along with our expansion for , (11.9), and (11.10) provide the expansion for as stated in Theorem 2.1.

    Last, since

    we can do similar computations to determine the expansion for .

    To get the expansions for the variables and , similar computations can be made, except with the constants , or instead of . The computations for these variables are nicer because , but one should use fifth-order expansions of and . So we have

    and are defined analogously. When one does the computations, one obtains the expansions for and listed in Theorem 2.1.

    Unlike the functions , and , which are in at time , the function has only been proven to be in at time , so the Taylor expansion can only go to fourth order. However, we still have which allows us to get constants in our expansion.

    The following equations are easy to compute from (2.5):

    (A.1a)
    (A.1b)
    (A.1c)
    (A.1d)
    (A.1e)
    (A.1f)
    (A.1g)
    (A.1h)
    (A.1i)
    (A.1j)
    (A.1k)
    (A.1l)
    (A.1m)
    (A.1n)

    If is a field, and denotes the field of formal Laurent series ¶¶ in the variable . The field of Puiseux series in the variable is then defined to be the union which is itself a field. The most important result concerning Puiseux series is the following:

    ¶¶Formal Laurent series are formal power series which allow for finitely many terms of negative degree, not to be confused with the Laurent series in complex analysis, which may have infinitely many terms of negative degree but must converge in an annulus.

    Theorem A.1 (Puiseux–Newton). If is an algebraically closed field of characteristic 0, then the field of Puiseux series with coefficients in an algebraically closed field. Furthermore, given a polynomial with , the coefficients of the roots of in can be constructed using the method of Newton polygons.

    Proof of Theorem A.1. See [27, Chapter Ⅳ, Section 3] or [28, Section 8.3].

    Of particular interest to us will be the following special case of the Puiseux–Newton theorem:

    Theorem A.2 (Analytic Puiseux–Newton). If denotes the ring of convergent power series in , and is a polynomial of degree , irreducible in , then there exists a convergent power series such that the roots of in are all given by

    It follows that in general if , then for each Puiseux series solution of , there exists some and such that .

    Proof of Theorem A.2. See [28, Section 8.3].

    Lemma A.3 (Quartic Inversion). There exists a constant and a nonempty open interval containing such that for all there exists a function defined for satisfying such that

    Furthermore, is an analytic function of satisfying the bounds

    for all , with the constant in the inequality independent of .

    Proof of Lemma A.3. The case where is trivial, so we will prove our result in the case . Define the recursive sequence ,

    and define the formal power series ,

    It is easy to check that is a Puiseux series solution to the algebraic equation . It follows from A.2 that must be convergent with some positive (possibly infinite) radius of convergence . Now pick any . If we define

    then it is easy to check that solves .

    Define the interval to be the range of , thought of as a function on , and define . Because everywhere, we know that for each , the equation has exactly one solution, . Therefore, if is a point such that , , and , then there exists with such that and since we conclude that and .

    The remaining expansion follows from the fact that and , combined with the fact that the power series is convergent.

    Theorem A.4. There exist universal constants and such that the following is true: Suppose that is an interval, , and is such that , , and has the Taylor expansion

    at . Then for all such that , we have

    where is a continuous function satisfying

    Proof of Theorem A.4. Assume without loss of generality that . We know that is away from and that

    for all . It follows from this formula that

    for all .

    First define the function ,

    Using our bounds on and , we see that

    Therefore, if we define and , then for all the function is strictly increasing and has a zero in the interior of . It follows from Corollary 1.1 in [29] that there exists a unique continuous function such that

    Now define the function ,

    It is easy to check that if and then

    Therefore, if and then for all the function is strictly increasing and contains a 0 in the interior of . It follows from Corollary 1.1 of [29] that there exists a unique continuous such that

    Our previous lemma A.3 tells us that there exist constants independent of or such that for all we have

    where . Now suppose that . Then and

    so . It follows that if then we have

    for all . Our result now follows.

    All authors contributed equally.

    The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.

    Steve Shkoller was supported in part by the NSF grant DMS-2007606, the Collaborative NSF grant DMS-2307680, and the Department of Energy Advanced Simulation and Computing (ASC) Program. Isaac Neal and Vlad Vicol were supported in part by the NSF CAREER grant DMS-1911413. Vlad Vicol was in part supported by the Collaborative NSF grant DMS-2307681, and a Simons Investigator Award.

    The authors declare there is no conflict of interest.



    [1] T. Buckmaster, T. D. Drivas, S. Shkoller, V. Vicol, Simultaneous development of shocks and cusps for 2D Euler with azimuthal symmetry from smooth data, Ann. PDE, 8 (2022). https://doi.org/10.1007/s40818-022-00141-6 doi: 10.1007/s40818-022-00141-6
    [2] T. Buckmaster, S. Shkoller, V. Vicol, Formation of shocks for 2D isentropic compressible Euler, Commun. Pur. Appl. Math., 75 (2022), 2069–2120. https://doi.org/10.1002/cpa.21956 doi: 10.1002/cpa.21956
    [3] P. D. Lax, Development of singularities of solutions of nonlinear hyperbolic partial differential equations, J. Math. Phys., 5 (1964), 611–613. https://doi.org/10.1063/1.1704154 doi: 10.1063/1.1704154
    [4] T. P. Liu, Development of singularities in the nonlinear waves for quasi-linear hyperbolic partial differential equations, J. Differ. Equations, 33 (1979), 92–111. https://doi.org/10.1016/0022-0396(79)90082-2 doi: 10.1016/0022-0396(79)90082-2
    [5] T. C. Sideris, Formation of singularities in three-dimensional compressible fluids, Commun. Math. Phys., 101 (1985), 475–485. https://doi.org/10.1007/BF01210741 doi: 10.1007/BF01210741
    [6] G. Chen, R. Pan, S. Zhu, Singularity formation for the compressible Euler equations, SIAM J. Math. Anal., 49 (2017), 2591–2614. https://doi.org/10.1137/16M1062818 doi: 10.1137/16M1062818
    [7] G. Chen, Optimal density lower bound on nonisentropic gas dynamics, J. Differ. Equations, 268 (2020), 4017–4028. https://doi.org/10.1016/j.jde.2019.10.017 doi: 10.1016/j.jde.2019.10.017
    [8] C. Dafermos, Hyperbolic Conservation Laws in Continuum Physics, Springer, 2005. https://doi.org/10.1007/978-3-662-49451-6
    [9] G. Q. Chen, D. Wang, The Cauchy problem for the Euler equations for compressible fluids, in Handbook of mathematical fluid dynamics, Vol. I (eds. S. Friedlander, and D. Serre), North-Holland, Amsterdam (2002), 421–543. https://doi.org/10.1016/s1874-5792(02)80012-x
    [10] T. P. Liu, Shock Waves, American Mathematical Society, Providence, RI, 2021.
    [11] J. Rauch, BV estimates fail for most quasilinear hyperbolic systems in dimensions greater than one, Commun. Math. Phys., 106 (1986), 481–484. https://doi.org/10.1007/BF01207258 doi: 10.1007/BF01207258
    [12] M. P. Lebaud, Description de la formation d'un choc dans le -système, Journal de mathématiques pures et appliquées, 73 (1994), 523–566.
    [13] S. Chen, L. Dong, Formation and construction of shock for -system, Sci. China Ser. A-Math., 44 (2001), 1139–1147. https://doi.org/10.1007/BF02877431 doi: 10.1007/BF02877431
    [14] D. X. Kong, Formation and propagation of singularities for quasilinear hyperbolic systems, Trans. Amer. Math. Soc., 354 (2002), 3155–3179. https://doi.org/10.1090/S0002-9947-02-02982-3 doi: 10.1090/S0002-9947-02-02982-3
    [15] T. Buckmaster, T. D. Drivas, S. Shkoller, V. Vicol, Formation and development of singularities for the compressible Euler equations, in Proceedings of the International Congress of Mathematicians 2022, EMS Press, Berlin, 5 (2023), 3636–3659. https://doi.org/10.4171/ICM2022/87
    [16] D. Christodoulou, The formation of shocks in 3-dimensional fluids, European Mathematical Society, 2007. https://doi.org/10.4171/031
    [17] D. Christodoulou, The shock development problem, European Mathematical Society, 2019. https://doi.org/10.4171/192
    [18] H. Yin, Formation and construction of a shock wave for 3-d compressible Euler equations with the spherical initial data, Nagoya Math. J., 175 (2004), 125–164. https://doi.org/10.1017/S002776300000893X doi: 10.1017/S002776300000893X
    [19] D. Christodoulou, A. Lisibach, Shock development in spherical symmetry, Ann. PDE, 2 (2016). https://doi.org/10.1007/s40818-016-0009-1 doi: 10.1007/s40818-016-0009-1
    [20] J. Luk, J. Speck, Shock formation in solutions to the 2d compressible Euler equations in the presence of non-zero vorticity, Invent. Mathe., 214 (2018), 1–169. https://doi.org/10.1007/s00222-018-0799-8 doi: 10.1007/s00222-018-0799-8
    [21] J. Luk, J. Speck, The stability of simple plane-symmetric shock formation for 3d compressible Euler flow with vorticity and entropy, Anal PDE, 17 (2024), 831–941. https://doi.org/10.2140/apde.2024.17.831 doi: 10.2140/apde.2024.17.831
    [22] T. Buckmaster, S. Shkoller, V. Vicol, Formation of point shocks for 3D compressible Euler, Comm. Pur. Appl. Math., 76 (2023), 2073–2191. https://doi.org/10.1002/cpa.22068 doi: 10.1002/cpa.22068
    [23] T. Buckmaster, S. Shkoller, V. Vicol, Shock formation and vorticity creation for 3D Euler, Comm. Pur. Appl. Math., 76 (2023), 1965–2072. https://doi.org/10.1002/cpa.22067 doi: 10.1002/cpa.22067
    [24] S. Shkoller, V. Vicol, The geometry of maximal development and shock formation for the Euler equations in multiple space dimensions, Invent. Math., 237 (2024), 871–1252. https://doi.org/10.1007/s00222-024-01269-x doi: 10.1007/s00222-024-01269-x
    [25] L. Abbrescia, J. Speck, The emergence of the singular boundary from the crease in compressible Euler flow, preprint, arXiv: 2207.07107.
    [26] L. D. Landau, E. M. Lifshitz, Fluid Mechanics, Vol. 6, Elsevier, 1987. https://doi.org/10.1016/C2013-0-03799-1
    [27] R. J. Walker, Algebraic Curves, Springer-Verlag, 1950.
    [28] E. Brieskorn, H. Knörrer, Plane algebraic curves, Springer Basel, 1986. https://doi.org/10.1007/978-3-0348-5097-1
    [29] S. Kumagai, An implicit function theorem: comment, J Optim Theory Appl, 31 (1980), 285–288. https://doi.org/10.1007/BF00934117 doi: 10.1007/BF00934117
  • Reader Comments
  • © 2025 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(201) PDF downloads(28) Cited by(0)

Article outline

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog