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

Local well-posedness of the Cauchy problem for a pp-adic Nagumo-type equation

L. F. Chacón-Cortés Pontificia Universidad Javeriana, Departamento de Matemáticas, Cra. 7 N. 40-62, Bogotá D.C., Colombia [email protected] C. A. Garcia-Bibiano Centro de Investigación y de Estudios Avanzados del Instituto Politécnico Nacional. Departamento de Matemáticas, Unidad Querétaro. Libramiento Norponiente #2000, Fracc. Real de Juriquilla. Santiago de Querétaro, Qro. 76230. México [email protected]  and  W. A. Zúñiga-Galindo1 University of Texas Rio Grande Valley. School of Mathematical & Statistical Sciences. One West University Blvd. Brownsville, TX 78520, United States [email protected]
Abstract.

We introduce a new family of pp-adic non-linear evolution equations. We establish the local well-posedness of the Cauchy problem for these equations in Sobolev-type spaces. For a certain subfamily, we show that the blow-up phenomenon occurs and provide numerical simulations showing this phenomenon.

Key words and phrases:
pp-adic analysis, pseudo-differential operators, Sobolev-type spaces, blow-up phenomenon.
2000 Mathematics Subject Classification:
Primary 47G30, 35B44; Secondary 46E36, 32P05
The third author was partially supported by the Lokenath Debnath Endowed Professorship, UTRGV

1. Introduction

Nowadays, the theory of linear partial pseudo-differential equations for complex-valued functions over pp-adic fields is a well-established branch of mathematical analysis, see e.g. [1]-[6], [12]-[16], [22]-[25], [27]-[33], and references therein. Meanwhile very little is known about nonlinear pp-adic equations. We can mention some semilinear evolution equations solved using pp-adic wavelets [1], [24], a kind of equations of reaction-diffusion type and Turing patterns studied in [31], [33], a pp-adic analog of one of the porous medium equation [17], [22], the blow-up phenomenon studied in [4], and non-linear integro-differential equations connected with pp-adic cellular networks [30].

In this article we introduce a new family of nonlinear evolution equations that we have named as pp-adic Nagumo-type equations:

ut=γ𝑫xαuu3+(β+1)u2βu+P(𝑫x)(um)xpn,t[0,T],u_{t}=-\gamma\boldsymbol{D}_{x}^{\alpha}u-u^{3}+\left(\beta+1\right)u^{2}-\beta u+P(\boldsymbol{D}_{x})\left(u^{m}\right)\text{, }x\in\mathbb{Q}_{p}^{n},\ t\in\left[0,T\right],

where γ>0\gamma>0, β0\beta\geq 0, 𝑫xα\boldsymbol{D}_{x}^{\alpha}, α>0\alpha>0, is the Taibleson operator, mm is a positive integer and P(𝑫x)P(\boldsymbol{D}_{x}) is an operator of degree δ\delta of the form P(𝑫)=j=0kCj𝑫δjP(\boldsymbol{D})=\sum_{j=0}^{k}C_{j}\boldsymbol{D}^{\delta_{j}}, where the CjC_{j}\in\mathbb{R} and δk=δ\delta_{k}=\delta. We establish the local well-posedness of the Cauchy problem for these equations in Sobolev-type spaces, see Theorem 1. For a certain subfamily, we show that the blow-up phenomenon occurs,  see Theorem 2, and we also provide numerical simulations showing this phenomenon.

The theory of Sobolev-type spaces use here was developed in [34], see also [25], [18]. This theory is based in the theory of countably Hilbert spaces of Gel’fand-Vilenkin [8]. Some generalizations are presented in [9]-[10]. We use classical techniques of operator semigroups, see e.g. [3], [20]. The family of evolution equations studied here contains as a particular case, equations of the form

(1.1) ut=γ𝑫xαuu3+(β+1)u2βu,u_{t}=-\gamma\boldsymbol{D}_{x}^{\alpha}u-u^{3}+\left(\beta+1\right)u^{2}-\beta u,

where xpn,t[0,T]x\in\mathbb{Q}_{p}^{n},\ t\in\left[0,T\right], 𝑫xα\boldsymbol{D}_{x}^{\alpha} is the Taibleson operator, that resemble the classical Nagumo-type equations, see e.g. [21].

In [7], the authors study the equations

(1.2) ut=Duxxu(uκ)(u1)εuxm,u_{t}=Du_{xx}-u\left(u-\kappa\right)\left(u-1\right)-\varepsilon u_{x}^{m},

where D>0D>0, κ(0,12)\kappa\in\left(0,\frac{1}{2}\right), ε>0\varepsilon>0, xx\in\mathbb{R}, t>0t>0. They establish the local well-posedness of the Cauchy problem for these equations in standard Sobolev spaces. There are several crucial differences between (1.1) and (1.2). The operators uxxu_{xx}, uxmu_{x}^{m} are local while the operators 𝑫xα\boldsymbol{D}_{x}^{\alpha}, P(𝑫x)(m)P(\boldsymbol{D}_{x})\left(\cdot^{m}\right) are non-local. The pp-adic heat equation ut=γ𝑫xαuu_{t}=-\gamma\boldsymbol{D}_{x}^{\alpha}u has an arbitrary order of pseudo-differentiability α>0\alpha>0 in the spatial variable, while in the classical fractional heat equation ut=Dμuxμu_{t}=D\frac{\partial^{\mu}u}{\partial x^{\mu}}, the degree of pseudo-differentiability μ(0,2]\mu\in\left(0,2\right]. This implies that the Markov processes attached to ut=γ𝑫xαuu_{t}=-\gamma\boldsymbol{D}_{x}^{\alpha}u are completely different to the ones attached to ut=Duxxu_{t}=Du_{xx}. In other words, the diffusion mechanisms in (1.1) and (1.2) are completely different. Notice that our non-linear term involves pseudo-derivatives of arbitrary order P(𝑫x)(um)P(\boldsymbol{D}_{x})\left(u^{m}\right), while in [7] only of first order uxmu_{x}^{m}.  Of course, the pp-adic Sobolev spaces behave completely different from their real counterparts, but the semigroup techniques are the same in both cases, since time is a non-negative real variable.

The article is organized as follows. In section 2, we review some basic aspects of the pp-adic analysis and fix the notation. In section 3, we present some technical results about Sobolev-type spaces and pp-adic pseudo-differential operators. In section 4, we show the local well-posedness of the pp-adic Nagumo-type equations, see Theorem 1. In section 5, we show a subfamily of pp-adic Nagumo-type equations whose solutions blow-up in finite time, see Theorem 2. In section 6, we present a numerical simulation showing the blow-up phenomenon.

2. pp-Adic Analysis: Essential Ideas

In this section, we collect some basic results on pp-adic analysis that we use through the article. For a detailed exposition the reader may consult [1], [14], [26], [29].

2.1. The field of pp-adic numbers

Along this article pp will denote a prime number. The field of pp-adic numbers p\mathbb{Q}_{p} is defined as the completion of the field of rational numbers \mathbb{Q} with respect to the pp-adic norm ||p|\cdot|_{p}, which is defined as

|x|p={0ifx=0pγifx=pγab,\left|x\right|_{p}=\left\{\begin{array}[c]{lll}0&\text{if}&x=0\\ &&\\ p^{-\gamma}&\text{if}&x=p^{\gamma}\frac{a}{b}\text{,}\end{array}\right.

where aa and bb are integers coprime with pp. The integer γ:=ord(x)\gamma:=ord(x), with ord(0):=+ord(0):=+\infty, is called the pp-adic order of xx.

Any pp-adic number x0x\neq 0 has a unique expansion of the form

x=pord(x)j=0xjpj,x=p^{ord(x)}\sum_{j=0}^{\infty}x_{j}p^{j},

where xj{0,,p1}x_{j}\in\{0,\dots,p-1\} and x00x_{0}\neq 0. By using this expansion, we define the fractional part of xpx\in\mathbb{Q}_{p}, denoted {x}p\{x\}_{p}, as the rational number

{x}p={0ifx=0 or ord(x)0pord(x)j=0ordp(x)1xjpjiford(x)<0.\left\{x\right\}_{p}=\left\{\begin{array}[c]{lll}0&\text{if}&x=0\text{ or }ord(x)\geq 0\\ &&\\ p^{ord(x)}\sum_{j=0}^{-ord_{p}(x)-1}x_{j}p^{j}&\text{if}&ord(x)<0.\end{array}\right.

2.2. Topology of pn\mathbb{Q}_{p}^{n}

For rr\in\mathbb{Z}, denote by Brn(a)={xpn;xappr}B_{r}^{n}(a)=\{x\in\mathbb{Q}_{p}^{n};||x-a||_{p}\leq p^{r}\} the ball of radius prp^{r} with center at a=(a1,,an)pna=(a_{1},\dots,a_{n})\in\mathbb{Q}_{p}^{n}, and take Brn(0):=BrnB_{r}^{n}(0):=B_{r}^{n}. Note that Brn(a)=Br(a1)××Br(an)B_{r}^{n}(a)=B_{r}(a_{1})\times\cdots\times B_{r}(a_{n}), where Br(ai):={xip;|xiai|ppr}B_{r}(a_{i}):=\{x_{i}\in\mathbb{Q}_{p};|x_{i}-a_{i}|_{p}\leq p^{r}\} is the one-dimensional ball of radius prp^{r} with center at aipa_{i}\in\mathbb{Q}_{p}. The ball B0nB_{0}^{n} equals the product of nn copies of B0=pB_{0}=\mathbb{Z}_{p}, the ring of pp-adic integers. We also denote by Srn(a)={xpn;xap=pr}S_{r}^{n}(a)=\{x\in\mathbb{Q}_{p}^{n};||x-a||_{p}=p^{r}\} the sphere of radius prp^{r} with center at a=(a1,,an)pna=(a_{1},\dots,a_{n})\in\mathbb{Q}_{p}^{n}, and take Srn(0):=SrnS_{r}^{n}(0):=S_{r}^{n}. We notice that S01=p×S_{0}^{1}=\mathbb{Z}_{p}^{\times} (the group of units of p\mathbb{Z}_{p}), but (p×)nS0n\left(\mathbb{Z}_{p}^{\times}\right)^{n}\subsetneq S_{0}^{n}. The balls and spheres are both open and closed subsets in pn\mathbb{Q}_{p}^{n}. In addition, two balls in pn\mathbb{Q}_{p}^{n} are either disjoint or one is contained in the other.

As a topological space (pn,||||p)\left(\mathbb{Q}_{p}^{n},||\cdot||_{p}\right) is totally disconnected, i.e. the only connected  subsets of pn\mathbb{Q}_{p}^{n} are the empty set and the points. A subset of pn\mathbb{Q}_{p}^{n} is compact if and only if it is closed and bounded in pn\mathbb{Q}_{p}^{n}, see e.g. [29, Section 1.3], or [1, Section 1.8]. The balls and spheres are compact subsets. Thus (pn,||||p)\left(\mathbb{Q}_{p}^{n},||\cdot||_{p}\right) is a locally compact topological space.

Since (pn,+)(\mathbb{Q}_{p}^{n},+) is a locally compact topological group, there exists a Haar measure dnxd^{n}x, which is invariant under translations, i.e. dn(x+a)=dnxd^{n}(x+a)=d^{n}x. If we normalize this measure by the condition pn𝑑x=1\int_{\mathbb{Z}_{p}^{n}}dx=1, then dnxd^{n}x is unique.

Notation 1.

We will use Ω(prxap)\Omega\left(p^{-r}||x-a||_{p}\right) to denote the characteristic function of the ball Brn(a)B_{r}^{n}(a). For more general sets, we will use the notation 1A1_{A} for the characteristic function of a set AA.

2.3. The Bruhat-Schwartz space

A complex-valued function φ\varphi defined on pn\mathbb{Q}_{p}^{n} is called locally constant if for any xpnx\in\mathbb{Q}_{p}^{n} there exist an integer l(x)l(x)\in\mathbb{Z} such that

(2.1) φ(x+x)=φ(x) for any xBl(x)n.\varphi(x+x^{\prime})=\varphi(x)\text{ for any }x^{\prime}\in B_{l(x)}^{n}.

A function φ:pn\varphi:\mathbb{Q}_{p}^{n}\rightarrow\mathbb{C} is called a Bruhat-Schwartz function (or a test function) if it is locally constant with compact support. Any test function can be represented as a linear combination, with complex coefficients, of characteristic functions of balls. The \mathbb{C}-vector space of Bruhat-Schwartz functions is denoted by 𝒟(pn):=𝒟\mathcal{D}(\mathbb{Q}_{p}^{n}):=\mathcal{D}. We denote by 𝒟(pn):=𝒟\mathcal{D}_{\mathbb{R}}(\mathbb{Q}_{p}^{n}):=\mathcal{D}_{\mathbb{R}} the \mathbb{R}-vector space of Bruhat-Schwartz functions. For φ𝒟(pn)\varphi\in\mathcal{D}(\mathbb{Q}_{p}^{n}), the largest number l=l(φ)l=l(\varphi) satisfying (2.1) is called the exponent of local constancy (or the parameter of constancy) of φ\varphi.

We denote by 𝒟ml(pn)\mathcal{D}_{m}^{l}(\mathbb{Q}_{p}^{n}) the finite-dimensional space of test functions from 𝒟(pn)\mathcal{D}(\mathbb{Q}_{p}^{n}) having supports in the ball BmnB_{m}^{n} and with parameters  of constancy l\geq l. We now define a topology on 𝒟\mathcal{D} as follows. We say that a sequence {φj}j\left\{\varphi_{j}\right\}_{j\in\mathbb{N}} of functions in 𝒟\mathcal{D} converges to zero, if the two following conditions hold:

(1) there are two fixed integers k0k_{0} and m0m_{0} such that  each φj\varphi_{j}\in 𝒟m0k0\mathcal{D}_{m_{0}}^{k_{0}};

(2) φj0\varphi_{j}\rightarrow 0 uniformly.

𝒟\mathcal{D} endowed with the above topology becomes a topological vector space.

2.4. LρL^{\rho} spaces

Given ρ[1,)\rho\in[1,\infty), we denote by Lρ:=Lρ(pn):=Lρ(pn,dnx),L^{\rho}:=L^{\rho}\left(\mathbb{Q}_{p}^{n}\right):=L^{\rho}\left(\mathbb{Q}_{p}^{n},d^{n}x\right), the \mathbb{C}-vector space of all the complex-valued functions gg satisfying

pn|g(x)|ρdnx<.{\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left|g\left(x\right)\right|^{\rho}d^{n}x<\infty.

The corresponding \mathbb{R}-vector spaces are denoted as Lρ:=Lρ(pn)=Lρ(pn,dnx)L_{\mathbb{R}}^{\rho}\allowbreak:=L_{\mathbb{R}}^{\rho}\left(\mathbb{Q}_{p}^{n}\right)=L_{\mathbb{R}}^{\rho}\left(\mathbb{Q}_{p}^{n},d^{n}x\right), 1ρ<1\leq\rho<\infty.

If UU is an open subset of pn\mathbb{Q}_{p}^{n}, 𝒟(U)\mathcal{D}(U) denotes the space of test functions with supports contained in UU, then 𝒟(U)\mathcal{D}(U) is dense in

Lρ(U)={φ:U;φρ={U|φ(x)|ρdnx}1ρ<},L^{\rho}\left(U\right)=\left\{\varphi:U\rightarrow\mathbb{C};\left\|\varphi\right\|_{\rho}=\left\{\int\limits_{U}\left|\varphi\left(x\right)\right|^{\rho}d^{n}x\right\}^{\frac{1}{\rho}}<\infty\right\},

where dnxd^{n}x is the normalized Haar measure on (pn,+)\left(\mathbb{Q}_{p}^{n},+\right), for 1ρ<1\leq\rho<\infty, see e.g. [1, Section 4.3]. We denote by Lρ(U)L_{\mathbb{R}}^{\rho}\left(U\right) the real counterpart of Lρ(U)L^{\rho}\left(U\right).

2.5. The Fourier transform

Set χp(y)=exp(2πi{y}p)\chi_{p}(y)=\exp(2\pi i\{y\}_{p}) for ypy\in\mathbb{Q}_{p}. The map χp()\chi_{p}(\cdot) is an additive character on p\mathbb{Q}_{p}, i.e. a continuous map from (p,+)\left(\mathbb{Q}_{p},+\right) into SS (the unit circle considered as multiplicative group) satisfying χp(x0+x1)=χp(x0)χp(x1)\chi_{p}(x_{0}+x_{1})=\chi_{p}(x_{0})\chi_{p}(x_{1}), x0,x1px_{0},x_{1}\in\mathbb{Q}_{p}.  The additive characters of p\mathbb{Q}_{p} form an Abelian group which is isomorphic to (p,+)\left(\mathbb{Q}_{p},+\right). The isomorphism is given by κχp(κx)\kappa\rightarrow\chi_{p}(\kappa x), see e.g. [1, Section 2.3].

Given ξ=(ξ1,,ξn)\xi=(\xi_{1},\dots,\xi_{n}) and y=(x1,,xn)pny=(x_{1},\dots,x_{n})\allowbreak\in\mathbb{Q}_{p}^{n}, we set ξx:=j=1nξjxj\xi\cdot x:=\sum_{j=1}^{n}\xi_{j}x_{j}. The Fourier transform of φ𝒟(pn)\varphi\in\mathcal{D}(\mathbb{Q}_{p}^{n}) is defined as

(φ)(ξ)=pnχp(ξx)φ(x)dnxfor ξpn,(\mathcal{F}\varphi)(\xi)={\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\chi_{p}(\xi\cdot x)\varphi(x)d^{n}x\quad\text{for }\xi\in\mathbb{Q}_{p}^{n},

where dnxd^{n}x is the normalized Haar measure on pn\mathbb{Q}_{p}^{n}. The Fourier transform is a linear isomorphism from 𝒟(pn)\mathcal{D}(\mathbb{Q}_{p}^{n}) onto itself satisfying

(2.2) ((φ))(ξ)=φ(ξ),(\mathcal{F}(\mathcal{F}\varphi))(\xi)=\varphi(-\xi),

see e.g. [1, Section 4.8]. We will also use the notation xξφ\mathcal{F}_{x\rightarrow\xi}\varphi and φ^\widehat{\varphi} for the Fourier transform of φ\varphi.

The Fourier transform extends to L2L^{2}. If fL2,f\in L^{2}, its Fourier transform is defined as

(f)(ξ)=limkxppkχp(ξx)f(x)dnx,for ξpn,(\mathcal{F}f)(\xi)=\lim_{k\rightarrow\infty}\int\limits_{||x||_{p}\leq p^{k}}\chi_{p}(\xi\cdot x)f(x)d^{n}x,\quad\text{for }\xi\in\mathbb{Q}_{p}^{n},

where the limit is taken in L2L^{2}. We recall that the Fourier transform is unitary on L2,L^{2}, i.e. f2=f2||f||_{2}=||\mathcal{F}f||_{2} for fL2f\in L^{2} and that (2.2) is also valid in L2L^{2}, see e.g. [26, Chapter III, Section 2].

2.6. Distributions

The \mathbb{C}-vector space 𝒟(pn)\mathcal{D}^{\prime}\left(\mathbb{Q}_{p}^{n}\right) :=𝒟:=\mathcal{D}^{\prime}  of all continuous linear functionals on 𝒟(pn)\mathcal{D}(\mathbb{Q}_{p}^{n}) is called the Bruhat-Schwartz space of distributions. Every linear functional on 𝒟\mathcal{D} is continuous, i.e. 𝒟\mathcal{D}^{\prime} agrees with the algebraic dual of 𝒟\mathcal{D}, see e.g. [29, Chapter 1, VI.3, Lemma]. We denote by 𝒟(pn)\mathcal{D}_{\mathbb{R}}^{\prime}\left(\mathbb{Q}_{p}^{n}\right) :=𝒟:=\mathcal{D}_{\mathbb{R}}^{\prime} the dual space of 𝒟\mathcal{D}_{\mathbb{R}}.

We endow 𝒟\mathcal{D}^{\prime} with the weak topology, i.e. a sequence {Tj}j𝕟\left\{T_{j}\right\}_{j\in\mathbb{n}} in 𝒟\mathcal{D}^{\prime} converges to TT if limjTj(φ)=T(φ)\lim_{j\rightarrow\infty}T_{j}\left(\varphi\right)=T\left(\varphi\right) for any φ𝒟\varphi\in\mathcal{D}.  The map

𝒟×𝒟(T,φ)T(φ)\begin{array}[c]{lll}\mathcal{D}^{\prime}\times\mathcal{D}&\rightarrow&\mathbb{C}\\ \left(T,\varphi\right)&\rightarrow&T\left(\varphi\right)\end{array}

is a bilinear form which is continuous in TT and φ\varphi separately. We call this map the pairing between 𝒟\mathcal{D}^{\prime} and 𝒟\mathcal{D}. From now on we will use (T,φ)\left(T,\varphi\right) instead of T(φ)T\left(\varphi\right).

Every ff in Lloc1L_{loc}^{1} defines a distribution f𝒟(pn)f\in\mathcal{D}^{\prime}\left(\mathbb{Q}_{p}^{n}\right) by the formula

(f,φ)=pnf(x)φ(x)dnx.\left(f,\varphi\right)={\textstyle\int\limits_{\mathbb{Q}_{p}^{n}}}f\left(x\right)\varphi\left(x\right)d^{n}x.

Such distributions are called regular distributions. Notice that for ff L2\in L_{\mathbb{R}}^{2}, (f,φ)=f,φ\left(f,\varphi\right)=\left\langle f,\varphi\right\rangle, where ,\left\langle\cdot,\cdot\right\rangle denotes the scalar product in L2L_{\mathbb{R}}^{2}.

2.7. The Fourier transform of a distribution

The Fourier transform [T]\mathcal{F}\left[T\right] of a distribution T𝒟(pn)T\in\mathcal{D}^{\prime}\left(\mathbb{Q}_{p}^{n}\right) is defined by

([T],φ)=(T,[φ]) for all φ𝒟(pn).\left(\mathcal{F}\left[T\right],\varphi\right)=\left(T,\mathcal{F}\left[\varphi\right]\right)\text{ for all }\varphi\in\mathcal{D}(\mathbb{Q}_{p}^{n})\text{.}

The Fourier transform T[T]T\rightarrow\mathcal{F}\left[T\right] is a linear (and continuous) isomorphism from 𝒟(pn)\mathcal{D}^{\prime}\left(\mathbb{Q}_{p}^{n}\right) onto 𝒟(pn)\mathcal{D}^{\prime}\left(\mathbb{Q}_{p}^{n}\right). Furthermore, T=[[T](ξ)]T=\mathcal{F}\left[\mathcal{F}\left[T\right]\left(-\xi\right)\right].

3. Sobolev-Type Spaces

The Sobolev-type spaces used here were introduce in [34], [25]. We follow here closely the presentation given in [18, Sections 10.1, 10.2].

We set [ξ]p:=max{1,ξp}\left[\xi\right]_{p}:=\max\left\{1,\left\|\xi\right\|_{p}\right\} for ξ=(ξ1,,ξn)pn\xi=(\xi_{1},\ldots,\xi_{n})\in\mathbb{Q}_{p}^{n}. Given φ,ϱ𝒟(pn)\varphi,\varrho\in\mathcal{D}(\mathbb{Q}_{p}^{n}) and ss\in\mathbb{R}, we define the scalar product:

φ,ϱs=pn[ξ]psφ^(ξ)ϱ^(ξ)¯dnξ,\langle\varphi,\varrho\rangle_{s}=\int\limits_{\mathbb{Q}_{p}^{n}}\left[\xi\right]_{p}^{s}\widehat{\varphi}(\xi)\overline{\widehat{\varrho}(\xi)}d^{n}\xi,

where the bar denotes the complex conjugate. We also set φs2=φ,φs\left\|\varphi\right\|_{s}^{2}=\langle\varphi,\varphi\rangle_{s}, and denote by s:=s(pn,)=s()\mathcal{H}_{s}:=\mathcal{H}_{s}(\mathbb{Q}_{p}^{n},\mathbb{C})=\mathcal{H}_{s}(\mathbb{C}) the completion of 𝒟(pn)\mathcal{D}(\mathbb{Q}_{p}^{n}) with respect to ,s\langle\cdot,\cdot\rangle_{s}. Notice that if r,sr,s\in\mathbb{R}, with rsr\leq s, then rs\left\|\cdot\right\|_{r}\leq\left\|\cdot\right\|_{s} and sr\mathcal{H}_{s}\hookrightarrow\mathcal{H}_{r} (continuous embedding). In particular,

21012,\cdots\supset\mathcal{H}_{-2}\supset\mathcal{H}_{-1}\supset\mathcal{H}_{0}\supset\mathcal{H}_{1}\supset\mathcal{H}_{2}\cdots,

where 0=L2\mathcal{H}_{0}=L^{2}. We set

(pn,)=:=ss.\mathcal{H}_{\infty}(\mathbb{Q}_{p}^{n},\mathbb{C})=\mathcal{H}_{\infty}:={\bigcap_{s\in\mathbb{N}}}\mathcal{H}_{s}.

Since [s]+1s[s]\mathcal{H}_{[s]+1}\subseteq\mathcal{H}_{s}\subseteq\mathcal{H}_{[s]} for s+s\in\mathbb{R}_{+}, where [][\cdot] is the integer part function, then =s+s\mathcal{H}_{\infty}={\bigcap_{s\in\mathbb{R}_{+}}}\mathcal{H}_{s}. With the topology induced by the family of seminorms {l}l\{\|\cdot\|_{l}\}_{l\in\mathbb{N}}, \mathcal{H}_{\infty} becomes a locally convex space, which is metrizable. Indeed,

d(f,g):=maxl{2lfgl1+fgl}, with f,g,d(f,g):=\max_{l\in\mathbb{N}}\left\{2^{-l}\frac{\|f-g\|_{l}}{1+\|f-g\|_{l}}\right\},\text{ with }f,g\in\mathcal{H}_{\infty},

is a metric for the topology of \mathcal{H}_{\infty} considered as a convex topological space. The metric space (,d)\left(\mathcal{H}_{\infty},d\right) is the completion of the metric space (𝒟(pn),d)(\mathcal{D}(\mathbb{Q}_{p}^{n}),d), cf. [18, Lemma 10.4]. Furthermore, LCunifL1L2\mathcal{H}_{\infty}\subset L^{\infty}\cap C^{\textup{unif}}\cap L^{1}\cap L^{2}, and (pn,)\mathcal{H}_{\infty}(\mathbb{Q}_{p}^{n},\mathbb{C}) is continuously embedded in C0(pn,)C_{0}(\mathbb{Q}_{p}^{n},\mathbb{C}). This is the non-Archimedean analog of the Sobolev embedding theorem, cf. [18, Theorem 10.15 ].

Lemma 1.

If s1ss2s_{1}\leq s\leq s_{2}, with s=θs1+(1θ)s2s=\theta s_{1}+(1-\theta)s_{2}, 0θ10\leq\theta\leq 1, then fsfs1θfs2(1θ)\left\|f\right\|_{s}\leq\left\|f\right\|_{s_{1}}^{\theta}\left\|f\right\|_{s_{2}}^{(1-\theta)}.

Proof.

Take fsf\in\mathcal{H}_{s}, then by using the Hölder inequality for the exponents 1q=θ,1q=1θ\frac{1}{q}=\theta,\frac{1}{q^{\prime}}=1-\theta,

fs2\displaystyle\left\|f\right\|_{s}^{2} =pn[ξ]ps|f^(ξ)|2dnξ=pn[ξ]pθs1+(1θ)s2|f^(ξ)|2(θ+(1θ))dnξ\displaystyle={\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left[\xi\right]_{p}^{s}\left|\widehat{f}\left(\xi\right)\right|^{2}d^{n}\xi={\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left[\xi\right]_{p}^{\theta s_{1}+(1-\theta)s_{2}}\left|\widehat{f}\left(\xi\right)\right|^{2\left(\theta+\left(1-\theta\right)\right)}d^{n}\xi
=pn([ξ]ps1|f^(ξ)|2)θ([ξ]ps2|f^(ξ)|2)1θdnξ\displaystyle={\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left(\left[\xi\right]_{p}^{s_{1}}\left|\widehat{f}\left(\xi\right)\right|^{2}\right)^{\theta}\left(\left[\xi\right]_{p}^{s_{2}}\left|\widehat{f}\left(\xi\right)\right|^{2}\right)^{1-\theta}d^{n}\xi
(pn[ξ]ps1|f^(ξ)|2dnξ)θ(pn[ξ]ps2|f^(ξ)|2dnξ)1θdnξ.\displaystyle\leq\left({\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left[\xi\right]_{p}^{s_{1}}\left|\widehat{f}\left(\xi\right)\right|^{2}d^{n}\xi\right)^{\theta}\left({\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left[\xi\right]_{p}^{s_{2}}\left|\widehat{f}\left(\xi\right)\right|^{2}d^{n}\xi\right)^{1-\theta}d^{n}\xi.

The following characterization of the spaces s\mathcal{H}_{s} and \mathcal{H}_{\infty} is useful:

Lemma 2 ([18, Lemma 10.8]).

(i) s={fL2;fs<}={T𝒟;Ts<}\mathcal{H}_{s}=\left\{f\in L^{2};\left\|f\right\|_{s}<\infty\right\}=\left\{T^{\prime}\in\mathcal{D};\left\|T\right\|_{s}<\infty\right\}, (ii) ={fL2;fs< for any s+}={T𝒟;Ts< for any s+}\mathcal{H}_{\infty}=\left\{f\in L^{2};\left\|f\right\|_{s}<\infty\text{ for any }s\in\mathbb{R}_{+}\right\}=\left\{T^{\prime}\in\mathcal{D};\left\|T\right\|_{s}<\infty\text{ for any }s\in\mathbb{R}_{+}\right\}. The equalities in (i)-(ii) are in the sense of vector spaces.

Proposition 1.

If s>n/2s>n/2, then s\mathcal{H}_{s} is a Banach algebra with respect to the product of functions. That is, if f,gsf,g\in\mathcal{H}_{s}, then fgsfg\in\mathcal{H}_{s} and fgsC(n,s)fsgs\left\|fg\right\|_{s}\leq C(n,s)\left\|f\right\|_{s}\left\|g\right\|_{s}, where C(n,s)C(n,s) is a positive constant.

Proof.

By the ultrametric property of p\left\|\cdot\right\|_{p}, ξpmax{ξηp,ηp}\left\|\xi\right\|_{p}\leq\max\left\{\left\|\xi-\eta\right\|_{p},\left\|\eta\right\|_{p}\right\} for ξ,ηpn\xi,\eta\in\mathbb{Q}_{p}^{n}, we have max{1,ξp}max{1,ξηp,ηp}\max\left\{1,\left\|\xi\right\|_{p}\right\}\leq\max\left\{1,\left\|\xi-\eta\right\|_{p},\left\|\eta\right\|_{p}\right\}, which implies that

[max{1,ξp}]smax{1,ξηps,ηps}=max{1,ξηp,ηp}s\left[\max\left\{1,\left\|\xi\right\|_{p}\right\}\right]^{s}\leq\max\left\{1,\left\|\xi-\eta\right\|_{p}^{s},\left\|\eta\right\|_{p}^{s}\right\}=\max\left\{1,\left\|\xi-\eta\right\|_{p},\left\|\eta\right\|_{p}\right\}^{s}

for s>0s>0. Therefore

(3.1) [ξ]ps[ξη]ps+[η]ps.\left[\xi\right]_{p}^{s}\leq\left[\xi-\eta\right]_{p}^{s}+\left[\eta\right]_{p}^{s}.

Now, for f,gL2f,g\in L^{2}, by using (3.1),

[ξ]ps2|fg^(ξ)|\displaystyle[\xi]_{p}^{\frac{s}{2}}\left|\widehat{fg}\left(\xi\right)\right| =|[ξ]ps2pnf^(ξη)g^(η)dnη|\displaystyle=\left|[\xi]_{p}^{\frac{s}{2}}{\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\widehat{f}(\xi-\eta)\widehat{g}(\eta)d^{n}\eta\right|
pn[ξη]ps2|f^(ξη)||g^(η)|dnη+pn[η]ps2|g^(η)||f^(ξη)|dnη\displaystyle\leq{\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}[\xi-\eta]_{p}^{\frac{s}{2}}\left|\widehat{f}(\xi-\eta)\right|\left|\widehat{g}(\eta)\right|d^{n}\eta+{\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}[\eta]_{p}^{\frac{s}{2}}\left|\widehat{g}(\eta)\right|\left|\widehat{f}(\xi-\eta)\right|d^{n}\eta
=[ξ]ps2|f^(ξ)||g^(ξ)|+|f^(ξ)|[ξ]ps2|g^(ξ)|.\displaystyle=[\xi]_{p}^{\frac{s}{2}}\left|\widehat{f}(\xi)\right|\ast\left|\widehat{g}(\xi)\right|+\left|\widehat{f}(\xi)\right|\ast[\xi]_{p}^{\frac{s}{2}}\left|\widehat{g}(\xi)\right|.

Then

fgs\displaystyle\left\|fg\right\|_{s} [ξ]ps2|f^(ξ)||g^(ξ)|+|f^(ξ)|[ξ]ps2|g^(ξ)|2\displaystyle\leq\left\|[\xi]_{p}^{\frac{s}{2}}\left|\widehat{f}(\xi)\right|\ast\left|\widehat{g}(\xi)\right|+\left|\widehat{f}(\xi)\right|\ast[\xi]_{p}^{\frac{s}{2}}\left|\widehat{g}(\xi)\right|\right\|_{2}
[ξ]ps2|f^(ξ)||g^(ξ)|2+|f^(ξ)|[ξ]ps2|g^(ξ)|2.\displaystyle\leq\left\|[\xi]_{p}^{\frac{s}{2}}\left|\widehat{f}(\xi)\right|\ast\left|\widehat{g}(\xi)\right|\right\|_{2}+\left\|\left|\widehat{f}(\xi)\right|\ast[\xi]_{p}^{\frac{s}{2}}\left|\widehat{g}(\xi)\right|\right\|_{2}.

Since [ξ]ps2|f^(ξ)|[\xi]_{p}^{\frac{s}{2}}\left|\widehat{f}(\xi)\right|, [ξ]ps2|g^(ξ)|L2[\xi]_{p}^{\frac{s}{2}}\left|\widehat{g}(\xi)\right|\in L^{2}, by using the Cauchy-Schwarz inequality with s>n/2s>n/2,  we have |g^(ξ)|1A(n,s)gs\left\|\left|\widehat{g}(\xi)\right|\right\|_{1}\leq A(n,s)\left\|g\right\|_{s}, |f^(ξ)|1A(n,s)fs\left\|\left|\widehat{f}(\xi)\right|\right\|_{1}\leq A(n,s)\left\|f\right\|_{s}, i.e. |g^(ξ)|\left|\widehat{g}(\xi)\right|, |f^(ξ)|L1\left|\widehat{f}(\xi)\right|\in L^{1}. Now, by the Young inequality, we obtain that

fgsfsg^1+gsf^12A(n,s)fsgs.\left\|fg\right\|_{s}\leq\left\|f\right\|_{s}\left\|\widehat{g}\right\|_{1}+\left\|g\right\|_{s}\left\|\widehat{f}\right\|_{1}\leq 2A(n,s)\left\|f\right\|_{s}\left\|g\right\|_{s}.

3.1. The Taibleson operator

Let α>0\alpha>0, the Taibleson operator is defined as

(𝑫αφ)(x)=ξx1(ξpα(xξφ)),(\boldsymbol{D}^{\alpha}\varphi)(x)=\mathcal{F}_{\xi\rightarrow x}^{-1}(\left\|\xi\right\|_{p}^{\alpha}(\mathcal{F}_{x\rightarrow\xi}\varphi)),

for φ𝒟(pn)\varphi\in\mathcal{D}(\mathbb{Q}_{p}^{n}). This operator admits the extension

(𝑫αf)(x)=1pα1pαnpnypαn{f(xy)f(x)}dny(\boldsymbol{D}^{\alpha}f)(x)=\frac{1-p^{\alpha}}{1-p^{-\alpha-n}}\int\limits_{\mathbb{Q}_{p}^{n}}\left\|y\right\|_{p}^{-\alpha-n}\{f(x-y)-f(x)\}d^{n}y

to locally constant functions satisfying

xp>1xpαn|f(x)|dnx<.\int\limits_{\left\|x\right\|_{p}>1}\left\|x\right\|_{p}^{-\alpha-n}\left|f\left(x\right)\right|d^{n}x<\infty.

The Taibleson operator 𝑫α\boldsymbol{D}^{\alpha} is the pp-adic analog of the fractional derivative. If n=1n=1, 𝑫α\boldsymbol{D}^{\alpha} agrees with the Vladimirov operator. The operator 𝑫α\boldsymbol{D}^{\alpha} does not satisfy the chain rule neither Leibniz formula. We also use the notation 𝑫xα\boldsymbol{D}_{x}^{\alpha}, when the Taibleson operator acts on functions depending on the variables xpnx\in\mathbb{Q}_{p}^{n} and t0t\geq 0.

Given 0=δ0<δ1<<δk1<δk=δ0=\delta_{0}<\delta_{1}<\cdots<\delta_{k-1}<\delta_{k}=\delta, we define

P(𝑫)=j=0kCj𝑫δj, where the Cj.P(\boldsymbol{D})={\displaystyle\sum\limits_{j=0}^{k}}C_{j}\boldsymbol{D}^{\delta_{j}}\text{, where the }C_{j}\in\mathbb{R}\text{.}
Lemma 3 ([18, Lemma 10.13 and Theorem 10.15]).

For s+s\in\mathbb{R}_{+}, the mapping P(𝐃):s+2δsP(\boldsymbol{D}):\mathcal{H}_{s+2\delta}\longrightarrow\mathcal{H}_{s} is a well-defined continuous mapping between Banach spaces.

Lemma 4.

Take s2δ>n/2s-2\delta>n/2 and f,gs+2δf,g\in\mathcal{H}_{s+2\delta}. Then

P(𝑫)(fg)sC(n,s,δ)fs+2δgs+2δ,\left\|P(\boldsymbol{D})\left(fg\right)\right\|_{s}\leq C(n,s,\delta)\left\|f\right\|_{s+2\delta}\left\|g\right\|_{s+2\delta},

where C(n,s,δ)C(n,s,\delta) is a positive constant that depends of nn, ss and δ\delta.

Proof.

Since s>n/2s>n/2 and f,gs+2δf,g\in\mathcal{H}_{s+2\delta}, by Proposition 1, fgs+2δfg\in\mathcal{H}_{s+2\delta}, and by Lemma 3, P(𝑫)(fg)sP(\boldsymbol{D})\left(fg\right)\in\mathcal{H}_{s}. Now by using Proposition 1,

P(𝑫)(fg)sj=0k|Cj|𝑫δj(fg)s\displaystyle\left\|P(\boldsymbol{D})\left(fg\right)\right\|_{s}\leq{\displaystyle\sum\limits_{j=0}^{k}}\left|C_{j}\right|\left\|\boldsymbol{D}^{\delta_{j}}\left(fg\right)\right\|_{s}
=j=0k|Cj|(pn[ξ]psξp2δj|fg^(ξ)|2dnξ)12j=0k|Cj|(pn[ξ]ps+2δj|fg^(ξ)|2dnξ)12\displaystyle={\displaystyle\sum\limits_{j=0}^{k}}\left|C_{j}\right|\left({\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left[\xi\right]_{p}^{s}\left\|\xi\right\|_{p}^{2\delta_{j}}\left|\widehat{fg}\left(\xi\right)\right|^{2}d^{n}\xi\right)^{\frac{1}{2}}\leq{\displaystyle\sum\limits_{j=0}^{k}}\left|C_{j}\right|\left({\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left[\xi\right]_{p}^{s+2\delta_{j}}\left|\widehat{fg}\left(\xi\right)\right|^{2}d^{n}\xi\right)^{\frac{1}{2}}
=j=0k|Cj|fgs+2δjj=0k|Cj|C(n,s,δj)fs+2δjgs+2δj\displaystyle={\displaystyle\sum\limits_{j=0}^{k}}\left|C_{j}\right|\left\|fg\right\|_{s+2\delta_{j}}\leq{\displaystyle\sum\limits_{j=0}^{k}}\left|C_{j}\right|C(n,s,\delta_{j})\left\|f\right\|_{s+2\delta_{j}}\left\|g\right\|_{s+2\delta_{j}}
(j=0k|Cj|C(n,s,δj))fs+2δgs+2δ.\displaystyle\leq\left({\displaystyle\sum\limits_{j=0}^{k}}\left|C_{j}\right|C(n,s,\delta_{j})\right)\left\|f\right\|_{s+2\delta}\left\|g\right\|_{s+2\delta}.

4. Local well-posedness of the pp-adic Nagumo-type equations

4.1. Some technical remarks

Let XX, YY Banach spaces, T0(0,)T_{0}\in(0,\infty) and let F:[0,T0]×YXF:\left[0,T_{0}\right]\times Y\longrightarrow X a continuous function. The Cauchy problem

(4.1) {tu(t)=F(t,u(t))u(0)=ϕY\left\{\begin{array}[c]{l}\partial_{t}u(t)=F\left(t,u\left(t\right)\right)\\ \\ u(0)=\phi\in Y\end{array}\right.

is locally well-posed in YY, if the following conditions are satisfied.

(i) There is T(0,T0]T\in\left(0,T_{0}\right] and a function uC([0,T];Y)u\in C([0,T];Y), with u(0)=ϕu(0)=\phi, satisfying the differential equation in the following sense:

limh0u(t+h)u(t)hF(t,u(t))X=0,\lim_{h\rightarrow 0}\left\|\frac{u(t+h)-u(t)}{h}-F(t,u(t))\right\|_{X}=0,

where the derivatives at t=0t=0 and t=Tt=T are calculated from the right and left, respectively.

(ii) The initial value problem (4.1) has at most one solution in C([0,T];Y)C([0,T];Y).

(iii) The function ϕu\phi\rightarrow u is continuous. That is, let {ϕn}\left\{\phi_{n}\right\} be a sequence in YY such that ϕnϕ\phi_{n}\rightarrow\phi_{\infty} in YY and let unC([0,Tn];Y)u_{n}\in C\left(\left[0,T_{n}\right];Y\right), resp. uC([0,T];Y)u_{\infty}\in C\left(\left[0,T_{\infty}\right];Y\right), be the corresponding solutions. Let T(0,T)T\in\left(0,T_{\infty}\right), then the solutions unu_{n} are defined in [0,T][0,T] for all nn big enough and

limnsupt[0,T]un(t)u(t)Y=0.\lim_{n\rightarrow\infty}\sup_{t\in[0,T]}\left\|u_{n}(t)-u_{\infty}(t)\right\|_{Y}=0.

4.2. Main result

Consider the following Cauchy problem:

(4.2) {uC([0,T],s)C1([0,T],s);ut=γ𝑫xαuu3+(β+1)u2βu+P(𝑫x)(um),xpn,t[0,T];u(0)=f0s,\left\{\begin{array}[c]{ll}u\in C\left(\left[0,T\right],\mathcal{H}_{s}\right)\cap C^{1}\left(\left[0,T\right],\mathcal{H}_{s}\right);&\\ &\\ u_{t}=-\gamma\boldsymbol{D}_{x}^{\alpha}u-u^{3}+\left(\beta+1\right)u^{2}-\beta u+P(\boldsymbol{D}_{x})\left(u^{m}\right),&x\in\mathbb{Q}_{p}^{n},\ t\in\left[0,T\right];\\ &\\ u(0)=f_{0}\in\mathcal{H}_{s},&\end{array}\right.

where TT, γ\gamma, α\alpha, β>0\beta>0, and mm is a positive integer. The main result of this work is the following:

Theorem 1.

For s>n/2+2δs>n/2+2\delta, the Cauchy problem (4.2) is locally well-posed in s\mathcal{H}_{s}.

4.3. Preliminary results

We denote by 𝑽(t)=e(γ𝑫α+β𝑰)t\boldsymbol{V}(t)=e^{-(\gamma\boldsymbol{D}^{\alpha}+\beta\boldsymbol{I})t}, t0t\geq 0, the semigroup in L2L^{2} generated by the operator 𝑨=γ𝑫αβ𝑰\boldsymbol{A}=-\gamma\boldsymbol{D}^{\alpha}-\beta\boldsymbol{I}, that is,

𝑽(t)f(x)=ξx1(e(γξpα+β)txξf), forfL2,t0.\boldsymbol{V}(t)f\left(x\right)=\mathcal{F}_{\xi\rightarrow x}^{-1}\left(e^{-(\gamma\lVert\xi\rVert_{p}^{\alpha}+\beta)t}\mathcal{F}_{x\rightarrow\xi}f\right)\text{, {for}}\ f\in L^{2},\ t\geq 0.
Lemma 5.

{𝑽(t)}t0\{\boldsymbol{V}(t)\}_{t\geq 0} is a C0C^{0}-semigroup of contractions in s\mathcal{H}_{s}, ss\in\mathbb{R}, satisfying 𝐕(t)seβt\left\|\boldsymbol{V}(t)\right\|_{s}\leq e^{-\beta t} for t0t\geq 0. Moreover, u(x,t)=𝐕(t)f0(x)u(x,t)=\boldsymbol{V}(t)f_{0}(x) is the unique solution to the following Cauchy problem:

(4.3) {uC([0,T],s)C1([0,T],s);ut=γ𝑫αuβu, t[0,T];u(x,0)=f0(x)s,\left\{\begin{array}[c]{l}u\in C\left(\left[0,T\right],\mathcal{H}_{s}\right)\cap C^{1}\left(\left[0,T\right],\mathcal{H}_{s}\right);\\ \\ u_{t}=-\gamma\boldsymbol{D}^{\alpha}u-\beta u,\text{ }t\in\left[0,T\right];\\ \\ u(x,0)=f_{0}(x)\in\mathcal{H}_{s},\end{array}\right.

where TT is an arbitrary positive number.

Proof.

We just verify the strongly continuity of the semigroup. Since

ξx1(e(γξpα+β)txξf)f(x)s2=pn[ξ]ps|f^(ξ)|2{1e(γξpα+β)t}2dnξfs2,\left\|\mathcal{F}_{\xi\rightarrow x}^{-1}\left(e^{-(\gamma\lVert\xi\rVert_{p}^{\alpha}+\beta)t}\mathcal{F}_{x\rightarrow\xi}f\right)-f\left(x\right)\right\|_{s}^{2}\\ ={\displaystyle\int\limits_{\mathbb{Q}_{p}^{n}}}\left[\xi\right]_{p}^{s}\left|\widehat{f}\left(\xi\right)\right|^{2}\left\{1-e^{-(\gamma\lVert\xi\rVert_{p}^{\alpha}+\beta)t}\right\}^{2}d^{n}\xi\leq\left\|f\right\|_{s}^{2},

it follows from the dominated convergence theorem that

limt0+𝑽(t)ffs=0.\lim_{t\rightarrow 0+}\left\|\boldsymbol{V}(t)f-f\right\|_{s}=0.

The existence and uniqueness of a solution for the Cauchy problem (4.3) follows from a well-known result, see e.g. [20, Theorem 4.3.1]. ∎

Lemma 6.

Let f0sf_{0}\in\mathcal{H}_{s}, ss\in\mathbb{R}, λ0\lambda\geq 0. Then, there exists a positive constant C(λ,α)C(\lambda,\alpha) that depends of λ\lambda and α\alpha such that

(4.4) 𝑽(t)f0s+λeβt(1+C(λ,α)(λ2αγt)λ2α)f0s for t>0.\lVert\boldsymbol{V}(t)f_{0}\rVert_{s+\lambda}\leq e^{-\beta t}\left(1+C(\lambda,\alpha)\left(\frac{\lambda}{2\alpha\gamma t}\right)^{\frac{\lambda}{2\alpha}}\right)\lVert f_{0}\rVert_{s}\text{ \ for }t>0\text{.}
Proof.

We first notice that

𝑽(t)f0s+λ2=pn[ξ]ps+λe2(γξpα+β)t|f0(ξ)|2dnξ\displaystyle\left\|\boldsymbol{V}\left(t\right)f_{0}\right\|_{s+\lambda}^{2}=\int\limits_{\mathbb{Q}_{p}^{n}}[\xi]_{p}^{s+\lambda}e^{-2(\gamma\lVert\xi\rVert_{p}^{\alpha}+\beta)t}\left|f_{0}\left(\xi\right)\right|^{2}d^{n}\xi
e2βt(supξpn[ξ]pλe2γξpαt)f0s2e2βt(1+supξpnpnξpλe2γξpαt)f0s2\displaystyle\leq e^{-2\beta t}\left(\sup_{\xi\in\mathbb{Q}_{p}^{n}}[\xi]_{p}^{\lambda}e^{-2\gamma\lVert\xi\rVert_{p}^{\alpha}t}\right)\left\|f_{0}\right\|_{s}^{2}\leq e^{-2\beta t}\left(1+\sup_{\xi\in\mathbb{Q}_{p}^{n}\smallsetminus\mathbb{Z}_{p}^{n}}\left\|\xi\right\|_{p}^{\lambda}e^{-2\gamma\lVert\xi\rVert_{p}^{\alpha}t}\right)\left\|f_{0}\right\|_{s}^{2}
e2βt(1+supξpnξpλe2γξpαt)f0s2.\displaystyle\leq e^{-2\beta t}\left(1+\sup_{\xi\in\mathbb{Q}_{p}^{n}}\left\|\xi\right\|_{p}^{\lambda}e^{-2\gamma\lVert\xi\rVert_{p}^{\alpha}t}\right)\left\|f_{0}\right\|_{s}^{2}.

We now set y=ξpy=\lVert\xi\rVert_{p} and h(y)=yλe2γyαth(y)=y^{\lambda}e^{-2\gamma y^{\alpha}t}. By using the fact that h(y)h(y) reaches its maximum at ymax=(λ2αγt)1αy_{\textup{max}}=\left(\frac{\lambda}{2\alpha\gamma t}\right)^{\frac{1}{\alpha}}, we conclude that

supξpnξpλe2γξpαt(λ2αγt)λαeλαC(λ,α)(λ2αγt)λα.\sup_{\xi\in\mathbb{Q}_{p}^{n}}\left\|\xi\right\|_{p}^{\lambda}e^{-2\gamma\lVert\xi\rVert_{p}^{\alpha}t}\leq\left(\frac{\lambda}{2\alpha\gamma t}\right)^{\frac{\lambda}{\alpha}}e^{-\frac{\lambda}{\alpha}}\leq\textup{C}\left(\lambda,\alpha\right)\left(\frac{\lambda}{2\alpha\gamma t}\right)^{\frac{\lambda}{\alpha}}.

Proposition 2.

Let s>n/2+2δs>n/2+2\delta and F(u)=(β+1)u2u3+P(𝐃)(um)F(u)=(\beta+1)u^{2}-u^{3}+P\left(\boldsymbol{D}\right)(u^{m}). Then F:ss2δF:\mathcal{H}_{s}\longrightarrow\mathcal{H}_{s-2\delta} is a continuous function satisfying

(4.5) F(u)F(w)s2δL(us,ws)uws,\lVert F(u)-F(w)\lVert_{s-2\delta}\leq L(\lVert u\lVert_{s},\lVert w\lVert_{s})\lVert u-w\lVert_{s},

for u,wsu,w\in\mathcal{H}_{s}, here L(,)L(\cdot,\cdot) is a continuous function, which is not decreasing with respect to each of their arguments. In particular,

(4.6) F(u)s2δL(us,0)us.\lVert F(u)\rVert_{s-2\delta}\leq L(\lVert u\rVert_{s},0)\lVert u\rVert_{s}.
Proof.

We first notice that

F(u)F(w)=(β+1)(u2w2)(u3w3)+P(𝑫)(umwm)\displaystyle F(u)-F(w)=(\beta+1)(u^{2}-w^{2})-(u^{3}-w^{3})+P\left(\boldsymbol{D}\right)(u^{m}-w^{m})
=(β+1)(uw)(u+w)(uw)(u2+uw+w2)+P(𝑫)((uw)q(u,w)),\displaystyle=(\beta+1)(u-w)\left(u+w\right)-(u-w)(u^{2}+uw+w^{2})+P\left(\boldsymbol{D}\right)((u-w)q(u,w)),

where q(u,w)=k=0m1ukwm1kq(u,w)=\sum_{k=0}^{m-1}u^{k}w^{m-1-k}. By using Proposition 1 and Lemma 4, the condition s>n/2s>n/2 implies that if u,wsu,w\in\mathcal{H}_{s}, then any polynomial function in u,wu,w belongs to s\mathcal{H}_{s}, and

F(u)F(w)s2δC{(β+1)uws2δu+ws2δ+uws2δu2+uw+w2s2δ+uwsq(u,w)s},\left\|F(u)-F(w)\right\|_{s-2\delta}\leq C\left\{(\beta+1)\lVert u-w\rVert_{s-2\delta}\lVert u+w\rVert_{s-2\delta}+\right.\\ \left.\lVert u-w\rVert_{s-2\delta}\lVert u^{2}+uw+w^{2}\rVert_{s-2\delta}+\lVert u-w\rVert_{s}\left\|q(u,w)\right\|_{s}\right\},

where C=C= C(n,s,δ)C(n,s,\delta). Then

F(u)F(w)s2δA(us,ws)uws,\left\|F(u)-F(w)\right\|_{s-2\delta}\leq A(\lVert u\rVert_{s},\lVert w\rVert_{s})\lVert u-w\lVert_{s},

where

A(us,ws)=C{(β+1)u+ws+u2+uw+w2s+q(u,w)s}C{(β+1)us+(β+1)ws+u2s+uws+w2s+k=0m1ukwm1ks}C(β+1)us+C(β+1)ws+C2us2+C2usws+C2ws2+Cm+1k=0m1uskwsm1k=:L(us,ws).A(\lVert u\rVert_{s},\lVert w\rVert_{s})=C\left\{(\beta+1)\lVert u+w\rVert_{s}+\lVert u^{2}+uw+w^{2}\rVert_{s}+\left\|q(u,w)\right\|_{s}\right\}\\ \leq C\left\{(\beta+1)\lVert u\rVert_{s}+(\beta+1)\lVert w\rVert_{s}+\lVert u^{2}\rVert_{s}+\lVert uw\rVert_{s}+\lVert w^{2}\rVert_{s}+\sum_{k=0}^{m-1}\left\|u^{k}w^{m-1-k}\right\|_{s}\right\}\\ \leq C(\beta+1)\lVert u\rVert_{s}+C(\beta+1)\lVert w\rVert_{s}+C^{2}\lVert u\rVert_{s}^{2}+C^{2}\lVert u\rVert_{s}\lVert w\rVert_{s}+C^{2}\lVert w\rVert_{s}^{2}+\\ C^{m+1}\sum_{k=0}^{m-1}\left\|u\right\|_{s}^{k}\left\|w\right\|_{s}^{m-1-k}=:L(\lVert u\lVert_{s},\lVert w\lVert_{s}).

For M,T>0M,T>0 and f0sf_{0}\in\mathcal{H}_{s}, we set

𝒳(M,T,f0):={u(t)C([0,T];s);supt[0,T]u(t)V(t)f0sM}.\mathcal{X}(M,T,f_{0}):=\left\{u(t)\in C\left([0,T];\mathcal{H}_{s}\right);\sup_{t\in[0,T]}\|u(t)-V(t)f_{0}\|_{s}\leq M\right\}.

We endow 𝒳(M,T,f0)\mathcal{X}(M,T,f_{0}) with the metric d(u(t),v(t))=supt[0,T]u(t)v(t)sd(u(t),v(t))=\sup_{t\in[0,T]}\|u(t)-v(t)\|_{s}. The resulting metric space is complete.

Proposition 3.

Take f0sf_{0}\in\mathcal{H}_{s} with s>n/2+2δs>n/2+2\delta, δ>0\delta>0. Then, there exists T=T(f0s,M)>0T=T(\lVert f_{0}\rVert_{s},M)>0 and a unique function uC([0,T];s)u\in C([0,T];\mathcal{H}_{s}) satisfying the integral equation

(4.7) u(t)=𝑽(t)f0+0t𝑽(tτ)F(u(τ))𝑑τ,u(t)=\boldsymbol{V}(t)f_{0}+\int\nolimits_{0}^{t}\boldsymbol{V}(t-\tau)F(u(\tau))d\tau,

such that u(0)=f0u(0)=f_{0}. Here F(u)=(β+1)u2u3+P(𝐃)(um)F(u)=(\beta+1)u^{2}-u^{3}+P(\boldsymbol{D})(u^{m}) as before.

Remark 1.

Since F(u)F(u) is not a locally Lipschitz function because inequality (4.6) involves two different norms, the existence of mild solutions of type (4.7) does not follow directly from standard results in semigroup theory, see e.g. [20, Theorem 5.2.2].

Proof.

Given u𝒳(M,T,f0)u\in\mathcal{X}(M,T,f_{0}), we set

𝑵u(t)=𝑽(t)f0+0t𝑽(tτ)F(u(τ))𝑑τ.\boldsymbol{N}u(t)=\boldsymbol{V}(t)f_{0}+\int\nolimits_{0}^{t}\boldsymbol{V}(t-\tau)F(u(\tau))d\tau.

Claim 1. 𝑵:𝒳(M,T,f0)C([0,T];s)\boldsymbol{N}:\mathcal{X}(M,T,f_{0})\longrightarrow C([0,T];\mathcal{H}_{s}).

Take u𝒳(M,T,f0)u\in\mathcal{X}(M,T,f_{0}), then

(4.8) 𝑵u(t1)𝑵u(t2)s(𝑽(t1)𝑽(t2))f0s\displaystyle\left\|\boldsymbol{N}u\left(t_{1}\right)-\boldsymbol{N}u\left(t_{2}\right)\right\|_{s}\leq\left\|\left(\boldsymbol{V}\left(t_{1}\right)-\boldsymbol{V}\left(t_{2}\right)\right)f_{0}\right\|_{s}
+0t1𝑽(t1τ)F(u(τ))𝑑τ0t2𝑽(t2τ)F(u(τ))𝑑τs.\displaystyle+\left\|\int\nolimits_{0}^{t_{1}}\boldsymbol{V}\left(t_{1}-\tau\right)F(u(\tau))d\tau-\int\nolimits_{0}^{t_{2}}\boldsymbol{V}\left(t_{2}-\tau\right)F(u(\tau))d\tau\right\|_{s}.

Since {𝑽(t)}t0\{\boldsymbol{V}(t)\}_{t\geq 0} is a C0C_{0}-semigroup in s\mathcal{H}_{s}, cf. Lemma 5, the first term on the right-hand side of the inequality (4.8) tends to zero when t2t1t_{2}\rightarrow t_{1}. To study the second term, we assume without loss of generality that 0<t1<t2<T0<t_{1}<t_{2}<T. Then

0t1𝑽(t1τ)F(u(τ))𝑑τ0t2𝑽(t2τ)F(u(τ))𝑑τs\displaystyle\left\|\int\nolimits_{0}^{t_{1}}\boldsymbol{V}\left(t_{1}-\tau\right)F(u(\tau))d\tau-\int\nolimits_{0}^{t_{2}}\boldsymbol{V}\left(t_{2}-\tau\right)F(u(\tau))d\tau\right\|_{s}
0t1{𝑽(t1τ)𝑽(t2τ)}F(u(τ))s𝑑τ+t1t2𝑽(t2τ)F(u(τ))s𝑑τ.\displaystyle\leq\int\nolimits_{0}^{t_{1}}\left\|\{\boldsymbol{V}\left(t_{1}-\tau\right)-\boldsymbol{V}\left(t_{2}-\tau\right)\}F(u(\tau))\right\|_{s}d\tau+\int\nolimits_{t_{1}}^{t_{2}}\left\|\boldsymbol{V}\left(t_{2}-\tau\right)F(u(\tau))\right\|_{s}d\tau.

By using Lemma 6 with λ=α\lambda=\alpha and Proposition 2,

(𝑽(t1τ)𝑽(t2τ))F(u(τ))s\displaystyle\left\|\left(\boldsymbol{V}\left(t_{1}-\tau\right)-\boldsymbol{V}\left(t_{2}-\tau\right)\right)F(u(\tau))\right\|_{s}
𝑽(t1τ)F(u(τ))s+𝑽(t2τ)F(u(τ))s\displaystyle\leq\left\|\boldsymbol{V}\left(t_{1}-\tau\right)F(u(\tau))\right\|_{s}+\left\|\boldsymbol{V}\left(t_{2}-\tau\right)F(u(\tau))\right\|_{s}
{2+C0(12γ(t1τ))12+C0(12γ(t2τ))12}F(u(τ))sα\displaystyle\leq\left\{2+C_{0}\left(\frac{1}{2\gamma(t_{1}-\tau)}\right)^{\frac{1}{2}}+C_{0}\left(\frac{1}{2\gamma(t_{2}-\tau)}\right)^{\frac{1}{2}}\right\}\|F(u(\tau))\|_{s-\alpha}
2{1+C0(12γ(t1τ))12}supτ[0,T]F(u(τ))sα\displaystyle\leq 2\left\{1+C_{0}\left(\frac{1}{2\gamma(t_{1}-\tau)}\right)^{\frac{1}{2}}\right\}\sup_{\tau\in[0,T]}\|F\left(u(\tau)\right)\|_{s-\alpha}
=A(T,s,α){1+C0(12γ(t1τ))12}L1([0,t1]).\displaystyle=A(T,s,\alpha)\left\{1+C_{0}\left(\frac{1}{2\gamma(t_{1}-\tau)}\right)^{\frac{1}{2}}\right\}\in L^{1}([0,t_{1}]).

Now, by applying the dominated convergence theorem,

limt2t10t1(𝑽(t1τ)𝑽(t2τ))F(u(τ))s𝑑τ=0.\lim\limits_{t_{2}\rightarrow t_{1}}\int\nolimits_{0}^{t_{1}}\left\|\left(\boldsymbol{V}\left(t_{1}-\tau\right)-\boldsymbol{V}\left(t_{2}-\tau\right)\right)F(u(\tau))\right\|_{s}d\tau=0.

By a similar argument, one shows that

𝑽(t2τ)F(u(τ))s2δ1+C0(12γ(t2τ))12L(u(τ)s,0)u(τ)s,\left\|\boldsymbol{V}\left(t_{2}-\tau\right)F(u(\tau))\right\|_{s-2\delta}\leq 1+C_{0}\left(\frac{1}{2\gamma(t_{2}-\tau)}\right)^{\frac{1}{2}}L(\lVert u\left(\tau\right)\rVert_{s},0)\lVert u\left(\tau\right)\rVert_{s},

and since

(4.9) u(τ)su(τ)𝑽(τ)f0s+𝑽(τ)f0sM+f0s, for all τ[0,T],\|u(\tau)\|_{s}\leq\|u(\tau)-\boldsymbol{V}(\tau)f_{0}\|_{s}+\|\boldsymbol{V}(\tau)f_{0}\|_{s}\leq M+\|f_{0}\|_{s},\text{ for all }\tau\in[0,T],

we have

(4.10) t1t2𝑽(t2τ)F(u(τ))s𝑑τ\displaystyle\int\nolimits_{t_{1}}^{t_{2}}\left\|\boldsymbol{V}(t_{2}-\tau)F(u(\tau))\right\|_{s}d\tau
L(M+f0s,0)(M+f0s)(t1t2(1+C0(12γ(t2τ))12)𝑑τ)\displaystyle\leq L(M+\|f_{0}\|_{s},0)(M+\|f_{0}\|_{s})\left(\int\nolimits_{t_{1}}^{t_{2}}\left(1+C_{0}\left(\frac{1}{2\gamma(t_{2}-\tau)}\right)^{\frac{1}{2}}\right)d\tau\right)
=L(M+f0s,0)(M+f0()s)((t2t1)+C0(2(t2t1)γ)),\displaystyle=L\left(M+\|f_{0}\|_{s},0\right)\left(M+\|f_{0}(\cdot)\|_{s}\right)\left((t_{2}-t_{1})+C_{0}\left(\sqrt{\frac{2(t_{2}-t_{1})}{\gamma}}\right)\right),

and consequently, by applying the dominated convergence theorem,

limt2t1t1t2𝑽(t2τ)F(u(τ))s𝑑τ=0.\lim_{t_{2}\rightarrow t_{1}}\int\nolimits_{t_{1}}^{t_{2}}\lVert\boldsymbol{V}(t_{2}-\tau)F(u(\tau))\rVert_{s}d\tau=0.

Claim 2. There exists T0T_{0} such that 𝑵(𝒳(M,T0,f0))𝒳(M,T0,f0)\boldsymbol{N}(\mathcal{X}(M,T_{0},f_{0}))\subseteq\mathcal{X}(M,T_{0},f_{0}).

By using a reasoning similar to the one used to established inequality (4.10), one gets

(𝑵u)(t)𝑽(t)f0s0t𝑽(tτ)F(u(τ))s𝑑τ\displaystyle\|(\boldsymbol{N}u)(t)-\boldsymbol{V}(t)f_{0}\|_{s}\leq\int\nolimits_{0}^{t}\|\boldsymbol{V}(t-\tau)F(u(\tau))\|_{s}d\tau
L(M+f0s,0)(M+f0s)(0t(1+C0(12γ(tτ))12)𝑑τ)\displaystyle\leq L\left(M+\|f_{0}\|_{s},0\right)\left(M+\|f_{0}\|_{s}\right)\left(\int\nolimits_{0}^{t}\left(1+C_{0}\left(\frac{1}{2\gamma(t-\tau)}\right)^{\frac{1}{2}}\right)d\tau\right)
L(M+f0s,0)(M+f0s)(T+C0(2Tγ)).\displaystyle\leq L\left(M+\|f_{0}\|_{s},0\right)\left(M+\|f_{0}\|_{s}\right)\left(T+C_{0}\left(\sqrt{\frac{2T}{\gamma}}\right)\right).

Now taking T0T_{0} such that

(4.11) L(M+f0s,0)(M+f0s)(T0+C0(2T0γ))M,L\left(M+\|f_{0}\|_{s},0\right)\left(M+\|f_{0}\|_{s}\right)\left(T_{0}+C_{0}\left(\sqrt{\frac{2T_{0}}{\gamma}}\right)\right)\leq M,

we conclude that 𝑵u𝒳(M,T0,f0)\boldsymbol{N}u\in\mathcal{X}(M,T_{0},f_{0}), for all u(t)𝒳(M,T0,f0)u(t)\in\mathcal{X}(M,T_{0},f_{0}).

Claim 3. There exists T0T_{0}^{\prime} such that 𝑵\boldsymbol{N}\ is a contraction on𝒳(M,T0,f0)\ \mathcal{X}(M,T_{0}^{\prime},f_{0}).

Given u(t)u(t), v(t)𝒳(M,T0,f0)v(t)\in\mathcal{X}(M,T_{0},f_{0}), by using Proposition 2, with

C0=L(M+f0s,M+f0s),C_{0}^{\prime}=L\left(M+\|f_{0}\|_{s},M+\|f_{0}\|_{s}\right),

see (4.9), we have

𝑵u(t)𝑵v(t)s\displaystyle\|\boldsymbol{N}u(t)-\boldsymbol{N}v(t)\|_{s} 0t𝑽(tτ)[F(u(τ))F(v(τ))]s𝑑τ\displaystyle\leq\int\nolimits_{0}^{t}\|\boldsymbol{V}(t-\tau)[F(u(\tau))-F(v(\tau))]\|_{s}d\tau
0t(1+C0(12γ(tτ))12)F(u(τ))F(v(τ))sα 𝑑τ\displaystyle\leq\int\nolimits_{0}^{t}\left(1+C_{0}\left(\frac{1}{2\gamma(t-\tau)}\right)^{\frac{1}{2}}\right)\|F(u(\tau))-F(v(\tau))\|_{s-\alpha}\text{ }d\tau
C00t(1+C0(12γ(tτ))12)u(τ)v(τ)s𝑑τ\displaystyle\leq C_{0}^{\prime}\int\nolimits_{0}^{t}\left(1+C_{0}\left(\frac{1}{2\gamma(t-\tau)}\right)^{\frac{1}{2}}\right)\|u(\tau)-v(\tau)\|_{s}d\tau
C0(supτ[0,T0]u(τ)v(τ)s)0t(1+C0(12γ(tτ))12)𝑑τ\displaystyle\leq C_{0}^{\prime}\left(\sup_{\tau\in[0,T_{0}]}\|u(\tau)-v(\tau)\|_{s}\right)\int\nolimits_{0}^{t}\left(1+C_{0}\left(\frac{1}{2\gamma(t-\tau)}\right)^{\frac{1}{2}}\right)d\tau
C0(T0+C0(2T0γ))d(u(t),v(t)).\displaystyle\leq C_{0}^{\prime}\left(T_{0}+C_{0}\left(\sqrt{\frac{2T_{0}}{\gamma}}\right)\right)d(u(t),v(t)).

Thus, taking T0T_{0}^{\prime} such that

(4.12) C:=C0(T0+C0(2T0γ))<1,C:=C_{0}^{\prime}\left(T_{0}^{\prime}+C_{0}\left(\sqrt{\frac{2T_{0}^{\prime}}{\gamma}}\right)\right)<1,

we obtain that d(𝑵u(t),𝑵v(t))Cd(u(t),v(t))d(\boldsymbol{N}u(t),\boldsymbol{N}v(t))\leq Cd(u(t),v(t)), that is, 𝑵\boldsymbol{N} is a strict contraction in 𝒳(M,T0,f0)\mathcal{X}(M,T_{0}^{\prime},f_{0}). We pick TT such that the inequalities (4.11) and (4.11) hold true, and apply the Banach Fixed Point Theorem to get u(t)𝒳(M,T,f0)u(t)\in\mathcal{X}(M,T,f_{0}) a unique fixed point of 𝑵\boldsymbol{N}, which satisfies the integral equation (4.7), where T=T(f0s,M)>0T=T(\lVert f_{0}\rVert_{s},M)>0. ∎

Remark 2.

Let 𝒳\mathcal{X} be a Banach space and let 𝐀:Dom(𝐀)𝒳\boldsymbol{A}:Dom(\boldsymbol{A})\rightarrow\mathcal{X} be an operator with dense domain such that 𝐀\boldsymbol{A} is the infinitesimal generator of a contraction semigroup (𝐒t)t0\left(\boldsymbol{S}_{t}\right)_{t\geq 0}. Fix T>0T>0 and let f:[0,T]𝒳f:\left[0,T\right]\rightarrow\mathcal{X} be a continuous function. Consider the Cauchy problem:

(4.13) {uC([0,T],Dom(𝑨))C1([0,T],𝒳);ut=𝑨u+f(t),t[0,T];u(0)=u0𝒳.\left\{\begin{array}[c]{l}u\in C\left(\left[0,T\right],Dom(\boldsymbol{A})\right)\cap C^{1}\left(\left[0,T\right],\mathcal{X}\right);\\ \\ u_{t}=\boldsymbol{A}u+f(t),\ \ t\in\left[0,T\right];\\ \\ u(0)=u_{0}\in\mathcal{X}.\end{array}\right.

Then

(4.14) u(t)=𝑺(t)u0+0t𝑺(tτ)f(τ))dτ,u(t)=\boldsymbol{S}(t)u_{0}+\int\nolimits_{0}^{t}\boldsymbol{S}(t-\tau)f(\tau))d\tau,

for t[0,T]t\in\left[0,T\right], see e.g. [3, Lemma 4.1.1]. Conversely, if u0Dom(𝐀)u_{0}\in Dom(\boldsymbol{A}), fC([0,T],𝒳)f\in C\left(\left[0,T\right],\mathcal{X}\right),

(0,T)f(τ)𝒳 𝑑τ<,{\displaystyle\int\limits_{\left(0,T\right)}}\left\|f\left(\tau\right)\right\|_{\mathcal{X}}\text{ }d\tau<\infty,

then a solution of (4.14) is a solution of the Cauchy problem (4.13), see e.g. [3, Proposition 4.1.6].

Proposition 4.

The problem (4.2) is equivalent to the integral equation (4.7). More precisely, if s>n/2+2δs>n/2+2\delta, and u(t)C([0,T];s)C1((0,T];s2δ)u(t)\in C([0,T];\mathcal{H}_{s})\cap C^{1}((0,T];\mathcal{H}_{s-2\delta}) is a solution of (4.2), then u(t)u(t) satisfies the integral equation (4.7). Conversely, if s>n/2+2δs>n/2+2\delta, and u(t)C([0,T];s)u(t)\in C([0,T];\mathcal{H}_{s}) is a solution of (4.7), then u(t)C1([0,T];s2δ)u(t)\in C^{1}([0,T];\mathcal{H}_{s-2\delta}) and it satisfies (4.2).

Proof.

It follows from Remark 2, Propositions 3, 2, by taking 𝑨=γ𝑫xαβ𝑰\boldsymbol{A=}-\gamma\boldsymbol{D}_{x}^{\alpha}-\beta\boldsymbol{I}, Dom(𝑨)=sDom(\boldsymbol{A})=\mathcal{H}_{s}, 𝒳=s2δ\mathcal{X}=\mathcal{H}_{s-2\delta}, f(t)=F(u(t))f(t)=F(u\left(t\right)). We first recall that 𝒟ss2δ\mathcal{D}\hookrightarrow\mathcal{H}_{s}\hookrightarrow\mathcal{H}_{s-2\delta}, where \hookrightarrow means continuous embedding,  an that 𝒟\mathcal{D}  is dense in s2δ\mathcal{H}_{s-2\delta}. If u(t)u(t) is a solution of (4.2), then, since F(u(t))C([0,T];s2δ)F(u\left(t\right))\in C([0,T];\mathcal{H}_{s-2\delta}), by Proposition 2, u(t)u\left(t\right) is a solution of (4.7). Conversely, if u(t)u\left(t\right) is a solution of (4.7), since

(0,T)F(u(τ))s2δ 𝑑τ<,{\displaystyle\int\limits_{\left(0,T\right)}}\left\|F(u\left(\tau\right))\right\|_{s-2\delta}\text{ }d\tau<\infty,

by Proposition 2, u(t)u\left(t\right) is a solution of (4.2). ∎

Lemma 7 ([20, Theorem 5.1.1]).

If hL1(0,T)h\in L^{1}\left(0,T\right), with T>0T>0, is real-valued function such that. If

h(t)a+b0th(s)𝑑s,h(t)\leq a+b\int\nolimits_{0}^{t}h(s)ds,

for tt\in (0,T)(0,T) a.e., where aa\in\mathbb{R} and b[0,)b\in\left[0,\infty\right) then h(t)aebth(t)\leq ae^{bt} for almost all tt in (0,T)(0,T).

Proposition 5.

Let f0f_{0}, f1sf_{1}\in\mathcal{H}_{s} and u(t),v(t)C[0,T];s)u(t),v(t)\in C[0,T];\mathcal{H}_{s}) be the corresponding solutions of equation (4.7) with initial conditions u(0)=f0u(0)=f_{0} and v(0)=f1v(0)=f_{1}, respectively. If s>n/2+2δs>n/2+2\delta, then

u(t)v(t)seL(W,W)f0f1s,\|u(t)-v(t)\|_{s}\leq e^{L\left(W,W\right)}\|f_{0}-f_{1}\|_{s},

where LL is given in Proposition 1 and

W:=max{supt[0,T]u(t)s,supt[0,T]v(t)s}.W:=\max\left\{\sup_{t\in[0,T]}\|u(t)\|_{s},\sup_{t\in[0,T]}\|v(t)\|_{s}\right\}.
Proof.

By using (4.7), we have

u(t)v(t)=𝑽(t)(f0f1)+0t𝑽(tτ){F(u(τ))F(v(τ))}𝑑τ.u(t)-v(t)=\boldsymbol{V}(t)(f_{0}-f_{1})+\int\nolimits_{0}^{t}\boldsymbol{V}(t-\tau)\{F(u(\tau))-F(v(\tau))\}d\tau.

By using Proposition 1, we get

u(t)v(t)sf0f1s+0t𝑽(tτ){F(u(τ))F(v(τ))}s𝑑τ\displaystyle\|u(t)-v(t)\|_{s}\leq\|f_{0}-f_{1}\|_{s}+\int\nolimits_{0}^{t}\|\boldsymbol{V}(t-\tau)\{F(u(\tau))-F(v(\tau))\}\|_{s}d\tau
f0f1s+0tF(u(τ))F(v(τ))sα 𝑑τ\displaystyle\leq\|f_{0}-f_{1}\|_{s}+\int\nolimits_{0}^{t}\|F(u(\tau))-F(v(\tau))\|_{s-\alpha}\text{ }d\tau
f0f1s+L(W,W)0tu(τ)v(τ)s𝑑τ.\displaystyle\leq\|f_{0}-f_{1}\|_{s}+L(W,W)\int\nolimits_{0}^{t}\|u(\tau)-v(\tau)\|_{s}d\tau.

Now the result follow from  Lemma 7, by taking  h(t)=u(t)v(t)sh(t)=\lVert u(t)-v(t)\rVert_{s}, a=f0f1sa=\lVert f_{0}-f_{1}\rVert_{s}, b=L(W,W)b=L(W,W). ∎

Proposition 6.

Let s>n/2+2δs>n/2+2\delta and δ0\delta\geq 0. Then, the map f0u(t)f_{0}\mapsto u(t) is continuous in the following sense: if f0(n)f0f_{0}^{(n)}\rightarrow f_{0} in s\mathcal{H}_{s} and un(t)C([0,Tn];s)u_{n}(t)\in C\left(\left[0,T_{n}\right];\mathcal{H}_{s}\right), with Tn=T(f0(n)s,M)>0T_{n}=T\left(\left\|f_{0}^{(n)}\right\|_{s},M\right)>0, are the corresponding solutions to the Cauchy problem (4.2) with un(0)=f0(n)u_{n}(0)=f_{0}^{(n)}. Then, there exist T>0T>0 and a positive integer N=N(γ,f0,T)N=N(\gamma,f_{0},T) such that TnTT_{n}\geq T for all nNn\geq N and

(4.15) limnsupt[0,T]un(t)u(t)s=0.\lim_{n\rightarrow\infty}\sup_{t\in[0,T]}\left\|u_{n}(t)-u(t)\right\|_{s}=0.
Proof.

By Proposition 3, the Tn=T(f0(n)s,M)>0T_{n}=T\left(\left\|f_{0}^{(n)}\right\|_{s},M\right)>0 are continuous functions of f0(n)s\left\|f_{0}^{(n)}\right\|_{s}, then, given T>0T^{\ast}>0 there exists NN\in\mathbb{N} such that TTnT^{\ast}\leq T_{n} for all nNn\geq N. We set T:=min{T,T1,T2,,TN1}T:=\min\left\{T^{\ast},T_{1},T_{2},\ldots,T_{N-1}\right\}. Therefore, all the un(t)u_{n}(t) are defined on [0,T][0,T], furthermore, u𝒳(M,T,f0(n))u\in\mathcal{X}\left(M,T,f_{0}^{(n)}\right) for all nn, and

un(t)sf0(n)s+Mδ+M,\left\|u_{n}(t)\right\|_{s}\leq\left\|f_{0}^{(n)}\right\|_{s}+M\leq\delta+M,

where δ=supnf0(n)s\delta=\sup_{n\in\mathbb{N}}\left\|f_{0}^{(n)}\right\|_{s}. Now

supt[0,T]un(t)sδ+M for all n,  andsupt[0,T]u(t)sδ+M.\sup_{t\in[0,T]}\left\|u_{n}(t)\right\|_{s}\leq\delta+M\text{ for all }n\text{, \ and}\sup_{t\in[0,T]}\|u(t)\|_{s}\leq\delta+M.

On the other hand, by reasoning as in the proof of Proposition 5, we have

un(t)u(t)sf0(n)f0s+L(δ+M,δ+M)0tun(τ)u(τ)s𝑑τ,\left\|u_{n}(t)-u(t)\right\|_{s}\leq\left\|f_{0}^{\left(n\right)}-f_{0}\right\|_{s}+L(\delta+M,\delta+M)\int\limits_{0}^{t}\|u_{n}(\tau)-u(\tau)\|_{s}d\tau,

and by applying Lemma 7

un(t)u(t)seTL(δ+M,δ+M)f0(n)f0s,\left\|u_{n}(t)-u(t)\right\|_{s}\leq e^{TL(\delta+M,\delta+M)}\left\|f_{0}^{\left(n\right)}-f_{0}\right\|_{s},

which in turns implies (4.15). ∎

4.4. Proof of the Main result

The local well-posedness of the Cauchy problem (4.2) in s\mathcal{H}_{s}, s>n/2+2δs>n/2+2\delta, follows from Propositions 3, 5, 6.

5. The Blow-up phenomenon

In this section, we study the blow-up phenomenon for the solution of the equation

(5.1) {ut=γ𝑫xαu+F(u)+𝑫xα1u3,xpn,t[0,T];u(0)=f0,\left\{\begin{array}[c]{ll}u_{t}=-\gamma\boldsymbol{D}_{x}^{\alpha}u+F(u)+\boldsymbol{D}_{x}^{\alpha_{1}}u^{3},&x\in\mathbb{Q}_{p}^{n},\ t\in\left[0,T\right];\\ &\\ u(0)=f_{0}\in\mathcal{H}_{\infty},&\end{array}\right.

where F(u)=u3+(β+1)u2βuF(u)=-u^{3}+\left(\beta+1\right)u^{2}-\beta u. We will say that a non-negative solution u(x,t)0u(x,t)\geq 0 of (5.1) blow-up in a finite time T>0T>0, if limtTsupxpnu(x,t)=+\lim_{t\rightarrow T^{-}}\sup_{x\in\mathbb{Q}_{p}^{n}}u(x,t)=+\infty. This limit makes sense since (pn,)\mathcal{H}_{\infty}(\mathbb{Q}_{p}^{n},\mathbb{C}) is continuously embedded in C0(pn,)C_{0}(\mathbb{Q}_{p}^{n},\mathbb{C}), [18, Theorem 10.15 ].

5.1. pp-adic wavelets and pseudo-differential operators

We denote by C(p,)C(\mathbb{Q}_{p},\mathbb{C}) the \mathbb{C}-vector space of  continuous \mathbb{C}-valued functions defined on p\mathbb{Q}_{p}.

We fix a function 𝔞:++\mathfrak{a}:\mathbb{R}_{+}\rightarrow\mathbb{R}_{+} and define the pseudo-differential operator

𝒟C(p,)L2φ𝑨φ,\begin{array}[c]{ccc}\mathcal{D}&\rightarrow&C(\mathbb{Q}_{p},\mathbb{C})\cap L^{2}\\ &&\\ \varphi&\rightarrow&\boldsymbol{A}\varphi,\end{array}

where (𝑨φ)(x)=ξx1{𝔞(|ξ|p)xξφ}\left(\boldsymbol{A}\varphi\right)\left(x\right)=\mathcal{F}_{\xi\rightarrow x}^{-1}\left\{\mathfrak{a}\left(\left|\xi\right|_{p}\right)\mathcal{F}_{x\rightarrow\xi}\varphi\right\}.

The set of functions {Ψrnj}\left\{\Psi_{rnj}\right\} defined as

(5.2) Ψrnj(x)=pr2χp(p1j(prxn))Ω(|prxn|p),\Psi_{rnj}\left(x\right)=p^{\frac{-r}{2}}\chi_{p}\left(p^{-1}j\left(p^{r}x-n\right)\right)\Omega\left(\left|p^{r}x-n\right|_{p}\right),

where rr\in\mathbb{Z}, j{1,,p1}j\in\left\{1,\cdots,p-1\right\}, and nn runs through a fixed set of representatives of p/p\mathbb{Q}_{p}/\mathbb{Z}_{p}, is an orthonormal basis of L2(p)L^{2}(\mathbb{Q}_{p}) consisting of eigenvectors of operator 𝑨\boldsymbol{A}:

(5.3) 𝑨Ψrnj=𝔞(p1r)Ψrnj for any rnj,\boldsymbol{A}\Psi_{rnj}=\mathfrak{a}(p^{1-r})\Psi_{rnj}\text{ for any }r\text{, }n\text{, }j\text{,}

see e.g. [18, Theorem 3.29], [1, Theorem 9.4.2]. Notice that

Ψ^rnj(ξ)=pr2χp(prnξ)Ω(|prξ+p1j|p),\widehat{\Psi}_{rnj}\left(\xi\right)=p^{\frac{r}{2}}\chi_{p}\left(p^{-r}n\xi\right)\Omega\left(\left|p^{-r}\xi+p^{-1}j\right|_{p}\right),

and then

𝔞(|ξ|p)Ψ^rnj(ξ)=𝔞(p1r)Ψ^rnj(ξ).\mathfrak{a}\left(\left|\xi\right|_{p}\right)\widehat{\Psi}_{rnj}\left(\xi\right)=\mathfrak{a}(p^{1-r})\widehat{\Psi}_{rnj}\left(\xi\right).

In particular, 𝑫xαΨrnj=p(1r)αΨrnj\boldsymbol{D}_{x}^{\alpha}\Psi_{rnj}=p^{\left(1-r\right)\alpha}\Psi_{rnj}, for any r,n,jr,n,j and α>0\alpha>0, and since p(1r)αp^{\left(1-r\right)\alpha},

𝑫xαRe(Ψrnj)=p(1r)αRe(Ψrnj)𝑫xαIm(Ψrnj)=p(1r)αIm(Ψrnj).\boldsymbol{D}_{x}^{\alpha}\operatorname{Re}\left(\Psi_{rnj}\right)=p^{\left(1-r\right)\alpha}\operatorname{Re}\left(\Psi_{rnj}\right)\text{, }\boldsymbol{D}_{x}^{\alpha}\operatorname{Im}\left(\Psi_{rnj}\right)=p^{\left(1-r\right)\alpha}\operatorname{Im}\left(\Psi_{rnj}\right).

And,

{Ψrn1(x)}2\displaystyle\left\{\Psi_{rn1}\left(x\right)\right\}^{2} =prχp(2p1(prxn))Ω(|prxn|p)\displaystyle=p^{-r}\chi_{p}\left(2p^{-1}\left(p^{r}x-n\right)\right)\Omega\left(\left|p^{r}x-n\right|_{p}\right)
=pr{Ψrn1(x)}2=pr2Ψrn2(x),\displaystyle=p^{r}\left\{\Psi_{rn1}\left(x\right)\right\}^{2}=p^{\frac{r}{2}}\Psi_{rn2}\left(x\right),

then

𝑫xαRe({Ψrn1(x)}2)=pr2p(1r)αRe(Ψrn2(x))=p(1r)αRe({Ψrn1(x)}2).\boldsymbol{D}_{x}^{\alpha}\operatorname{Re}\left(\left\{\Psi_{rn1}\left(x\right)\right\}^{2}\right)=p^{\frac{r}{2}}p^{\left(1-r\right)\alpha}\operatorname{Re}\left(\Psi_{rn2}(x)\right)=p^{\left(1-r\right)\alpha}\operatorname{Re}\left(\left\{\Psi_{rn1}\left(x\right)\right\}^{2}\right).

5.2. The blow-up

In this section, we assume that u(x,t)u(x,t) is real-valued non-negative solution of the Cauchy problem (4.2) in \mathcal{H}_{\infty}. We set w(x):=Re({Ψrn1(x)}2)w(x):=\operatorname{Re}\left(\left\{\Psi_{rn1}\left(x\right)\right\}^{2}\right), so 𝑫xαw(x)=p(1r)αw(x)\boldsymbol{D}_{x}^{\alpha}w(x)=p^{\left(1-r\right)\alpha}w(x). Thus w(x)dxw(x)dx defines a (positive) measure. We also set G(t):=pu(x,t)w(x)𝑑xG(t):=\int_{\mathbb{Q}_{p}}u(x,t)w(x)dx, where u(x,t)u(x,t) is a positive solution of (5.1), then

G(t)=put(x,t)w(x)𝑑x=γp(𝑫xαu)(x,t)w(x)𝑑x\displaystyle G^{\prime}(t)=\int\limits_{\mathbb{Q}_{p}}u_{t}(x,t)w(x)dx=-\gamma\int\limits_{\mathbb{Q}_{p}}(\boldsymbol{D}_{x}^{\alpha}u)(x,t)w(x)dx
(5.4) +pF(u(x,t))w(x)𝑑x+p(𝑫xα1u3)(x,t)w(x)𝑑x.\displaystyle+\int\limits_{\mathbb{Q}_{p}}F(u(x,t))w(x)dx+\int\limits_{\mathbb{Q}_{p}}(\boldsymbol{D}_{x}^{\alpha_{1}}u^{3})(x,t)w(x)dx.

Now, by using that 𝑫xαu(,t)\boldsymbol{D}_{x}^{\alpha}u(\cdot,t), wL2w\in L^{2}, and F(u(,t))F(u(\cdot,t)), 𝑫xα1u3(,t)L2\boldsymbol{D}_{x}^{\alpha_{1}}u^{3}(\cdot,t)\in L^{2} since for s>n/2s>n/2, s\mathcal{H}_{s} is a Banach algebra contained in L2L^{2} cf. Proposition 1, and applying the Parseval-Steklov theorem, we get (5.4) can be rewritten as

G(t)=p(γp(1r)αu(x,t)+F(u(x,t))+p(1r)α1u3(x,t))w(x)𝑑x.G^{\prime}(t)=\int\limits_{\mathbb{Q}_{p}}\left(-\gamma p^{\left(1-r\right)\alpha}u(x,t)+F(u(x,t))+p^{\left(1-r\right)\alpha_{1}}u^{3}(x,t)\right)w(x)dx.

Since the function H(y)=γp(1r)αy+F(y)+p(1r)α1y3H(y)=-\gamma p^{\left(1-r\right)\alpha}y+F(y)+p^{\left(1-r\right)\alpha_{1}}y^{3} is convex because

H′′(y)=6y+2(β+1)+p(1r)α16y=6y(p(1r)α11)+2(β+1)0,H^{\prime\prime}(y)=-6y+2\left(\beta+1\right)+p^{\left(1-r\right)\alpha_{1}}6y=6y\left(p^{\left(1-r\right)\alpha_{1}}-1\right)+2\left(\beta+1\right)\geq 0,

for y0y\geq 0, and r0r\leq 0, we can use the Jensen’s inequality to get G(t)H(G(t))G^{\prime}(t)\geq H(G(t)), then the function G(t)G(t) can not remain finite for every t[0,)t\in[0,\infty). Then there exists T(0,)T\in(0,\infty) such that limtTG(t)=+\lim_{t\rightarrow T^{-}}G(t)=+\infty, hence u(x,t)u(x,t) blow ups at the time TT. Then we have established the following result:

Theorem 2.

Let u(x,t)u(x,t) be a positive solution of (5.1). Then there T(0,+)T\in\left(0,+\infty\right) depending on the initial datum such that limtTsupxpnu(x,t)=+\lim_{t\rightarrow T^{-}}\sup_{x\in\mathbb{Q}_{p}^{n}}u(x,t)=+\infty.

6. Numerical Simulations

In this section, we present two numerical simulations for the solution of problem (5.1) (in dimension one) for a suitable initial datum. We solve and visualize (using a heat map) the radial profiles of the solution of (5.1). We consider equation (5.1) for radial functions u(x,)u(x,\cdot). In [15], Kochubei obtained a formula for 𝑫xαu(x,t)\boldsymbol{D}_{x}^{\alpha}u\left(x,t\right) as an absolutely convergent real series, we truncate this series and then we apply the classic Euler Forward Method (see e.g. [23]) to find the values of u(pord(x),t)u(p^{-ord(x)},t), when 20ord(x)20-20\leq ord(x)\leq 20 (vertical axis) and when t={tk:tk=1/k,k=1,,300}t=\{t_{k}:\,\,t_{k}=1/k,k=1,\dots,300\} (horizontal axis).  In Figure 1, on the left, the heat map of the numerical solution of the homogeneous equation ut(x,t)=𝑫xαu(x,t)u_{t}(x,t)=-\boldsymbol{D}_{x}^{\alpha}u(x,t) with initial data u(x,0)=4ep|ord(x)|/100u(x,0)=4e^{-p^{|ord(x)|}/100} (Gaussian bell type), and parameters p=3p=3, α=0.2\alpha=0.2, γ=1\gamma=1. On the right side, we have the numerical solution of the equation ut(x,t)=𝑫xαu(x,t)u3(x,t)+(β+1)u2(x,t)βu(x,t)+𝑫xα1u3(x,t)u_{t}(x,t)=-\boldsymbol{D}_{x}^{\alpha}u(x,t)-u^{3}(x,t)+(\beta+1)u^{2}(x,t)-\beta u(x,t)+\boldsymbol{D}_{x}^{\alpha_{1}}u^{3}(x,t), with p=3p=3, α=0.2\alpha=0.2, α1=0.1\alpha_{1}=0.1, and β=0.7\beta=0.7.

On the left side of the Figure 1, we observe that the solution uu is uniformly decreasing with respect to the variable tt. This behavior is typical for solutions of diffusion equations. These equations have been extensively studied, see e.g. [18], [35] and the references therein.

On the right side of Figure 1, we see that the evolution of u(x,t)u(x,t) is controlled by the diffusion term 𝑫xαu(x,t)-\boldsymbol{D}_{x}^{\alpha}u(x,t), up to a time TT (blow-up time), this behavior is similar to that described above. When t>Tt>T, the reactive term u3(x,t)+(β+1)u2(x,t)βu(x,t)+𝑫xα1u3(x,t)-u^{3}(x,t)+(\beta+1)u^{2}(x,t)-\beta u(x,t)+\boldsymbol{D}_{x}^{\alpha_{1}}u^{3}(x,t) takes over and u(x,t)u(x,t) grows rapidly towards infinity.

The method converges quite fast, but still lacks a mathematical formalism to support it, for this reason we refer to it as a numerical simulation of the solution.

References

  • [1] Albeverio S., Khrennikov A. Yu., Shelkovich V. M., Theory of pp-adic distributions: linear and nonlinear models. London Mathematical Society Lecture Note Series, 370. Cambridge University Press, Cambridge, 2010.
  • [2] Albeverio S., Khrennikov A. Yu., Shelkovich V. M., The Cauchy problems for evolutionary pseudo-differential equations over pp-adic field and the wavelet theory, J. Math. Anal. Appl. 375 (2011), no. 1, 82–98.
  • [3] Cazenave Thierry, Haraux Alain, An introduction to semilinear evolution equations. Oxford University Press, 1998.
  • [4] Chacón-Cortés L. F., Gutiérrez-García Ismael, Torresblanca-Badillo Anselmo, Vargas Andrés Finite time blow-up for a p-adic nonlocal semilinear ultradiffusion equation, J. Math. Anal. Appl. 494 (2021), no. 2, Paper No. 124599, 22 pp.
  • [5] Chacón-Cortés L. F., Zúñiga-Galindo W. A., Non-local operators, non-Archimedean parabolic-type equations with variable coefficients and Markov processes, Publ. Res. Inst. Math. Sci. 51 (2015), no. 2, 289–317.
  • [6] Chacón-Cortés L. F., Zúñiga-Galindo W. A. Nonlocal operators, parabolic-type equations, and ultrametric random walks, J. Math. Phys. 54 (2013), no. 11, 113503, 17 pp.
  • [7] De la Cruz Richard, Lizarazo Vladimir , Local well-posedness to the Cauchy problem for an equation of Nagumo type. Preprint 2019.
  • [8] Gel’fand I.M., Vilenkin N.Y., Generalized Functions. Applications of Harmonic Analysis, vol. 4. Academic Press, New York, 1964.
  • [9] Górka Przemysław, Kostrzewa Tomasz, Reyes Enrique G., Sobolev spaces on locally compact abelian groups: compact embeddings and local spaces, J. Funct. Spaces 2014, Art. ID 404738, 6 pp.
  • [10] Górka Przemysław, Kostrzewa Tomasz, Sobolev spaces on metrizable groups, Ann. Acad. Sci. Fenn. Math. 40 (2015), no. 2, 837–849.
  • [11] Halmos Paul R., Measure Theory. D. Van Nostrand Co., Inc., New York, N.Y., 1950.
  • [12] Haran S., Quantizations and symbolic calculus over the pp-adic numbers. Ann. Inst. Fourier 43 (1993), no. 4, 997–1053.
  • [13] Kaneko H., Besov space and trace theorem on a local field and its application, Math. Nachr. 285 (2012), no. 8-9, 981–996.
  • [14] Kochubei A.N., Pseudo-Differential Equations and Stochastics over Non-Archimedean Fields. Marcel Dekker, New York, 2001.
  • [15] Kochubei A. N., Radial solutions of non-Archimedean pseudodifferential equations, Pacific J. Math. 269 (2014), no. 2, 355–369.
  • [16] Kochubei A.N., A non-Archimedean wave equation, Pacific J. Math. 235 (2008), no. 2, 245–261.
  • [17] Khrennikov Andrei Yu, Kochubei Anatoly N., pp-Adic Analogue of the Porous Medium Equation, J Fourier Anal Appl (2018) 24:1401–1424.
  • [18] Khrennikov Andrei Yu., Kozyrev Sergei V., Zúñiga-Galindo W. A., Ultrametric pseudodifferential equations and applications. Encyclopedia of Mathematics and its Applications, 168. Cambridge University Press, Cambridge, 2018.
  • [19] Khrennikov Andrei, Oleschko Klaudia, Correa López,Maria de Jesús, Application of p-adic wavelets to model reaction-diffusion dynamics in random porous media, J. Fourier Anal. Appl. 22 (2016), no. 4, 809–822.
  • [20] Miklavčič Milan. Applied functional analysis and partial differential equations. World Scientific Publishing Co., Inc., River Edge, NJ, 1998.
  • [21] Nagumo J., Yoshizawa S. and Arimoto S. Bistable Transmission Lines. IEEE Transactions on Circuit Theory, vol. 12, no. 3, pp. 400-412, September 1965.
  • [22] Oleschko K., Khrennikov A., Transport through a network of capillaries from ultrametric diffusion equation with quadratic nonlinearity, Russ. J. Math. Phys. 24 (2017), no. 4, 505–516.
  • [23] Press W. H., Flannery B. P., Teukolsky, S. A., and Vetterling W. T., Numerical Recipes in FORTRAN: The Art of Scientific Computing, 2nd ed. Cambridge, England: Cambridge University Press, p. 710, 1992.
  • [24] Pourhadi Ehsan, Khrennikov Andrei Yu., Oleschko Klaudia, Correa Lopez María de Jesús, Solving nonlinear pp-adic pseudo-differential equations: combining the wavelet basis with the Schauder fixed point theorem, J. Fourier Anal. Appl. 26 (2020), no. 4, Paper No. 70, 23 pp.
  • [25] Rodríguez-Vega J. J., Zúñiga-Galindo W. A., Elliptic pseudodifferential equations and Sobolev spaces over pp-adic fields, Pacific J. Math. 246 (2010), no. 2, 407–420.
  • [26] Taibleson M.H., Fourier Analysis on Local Fields. Princeton University Press, Princeton, 1975.
  • [27] Torresblanca-Badillo Anselmo, Zúñiga-Galindo W. A., Ultrametric diffusion, exponential landscapes, and the first passage time problem, Acta Appl. Math. 157 (2018), 93–116.
  • [28] Torresblanca-Badillo Anselmo, Zúñiga-Galindo W. A., Non-Archimedean pseudodifferential operators and Feller semigroups, pp-Adic Numbers Ultrametric Anal. Appl. 10 (2018), no. 1, 57–73.
  • [29] Vladimirov V. S., Volovich I. V. and Zelenov E. I., pp-adic analysis and mathematical physics, World Scientific, 1994.
  • [30] Zambrano-Luna B., Zúñiga-Galindo W. A., pp-Adic Cellular Neural Networks. https://arxiv.org/abs/2107.07980. 
  • [31] Zúñiga-Galindo W. A., Reaction-diffusion equations on complex networks and Turing patterns, via pp-adic analysis, J. Math. Anal. Appl. 491 (2020), no. 1, 124239, 39 pp.
  • [32] Zúñiga-Galindo W. A., Non-archimedean replicator dynamics and Eigen’s paradox, J. Phys. A 51 (2018), no. 50, 505601, 26 pp.
  • [33] Zúñiga-Galindo W. A., Non-Archimedean reaction-ultradiffusion equations and complex hierarchic systems, Nonlinearity 31 (2018), no. 6, 2590–2616.
  • [34] Zúñiga-Galindo W. A. Non-Archimedean white noise, pseudodifferential stochastic equations, and massive Euclidean fields, J. Fourier Anal. Appl. 23 (2017), no. 2, 288–323.
  • [35] Zúñiga-Galindo W. A., Pseudodifferential equations over non-Archimedean spaces. Lecture Notes in Mathematics, 2174. Springer, Cham, 2016.
  • [36] Zúñiga-Galindo W. A., The Cauchy problem for non-Archimedean pseudodifferential equations of Klein-Gordon type, J. Math. Anal. Appl. 420 (2014), no. 2, 1033–1050.
  • [37] Zúñiga-Galindo W. A., Parabolic equations and Markov processes over pp-adic fields. Potential Anal. 28 (2008), no. 2, 185–200.
  • [38] Zuniga-Galindo W. A., Fundamental solutions of pseudo-differential operators over pp-adic fields. Rend, Sem. Mat. Univ. Padova 109 (2003), 241–245.