Izvestiya: Mathematics
RUS  ENG    JOURNALS   PEOPLE   ORGANISATIONS   CONFERENCES   SEMINARS   VIDEO LIBRARY   PACKAGE AMSBIB  
General information
Latest issue
Forthcoming papers
Archive
Impact factor
Guidelines for authors
Submit a manuscript

Search papers
Search references

RSS
Latest issue
Current issues
Archive issues
What is RSS



Izv. RAN. Ser. Mat.:
Year:
Volume:
Issue:
Page:
Find






Personal entry:
Login:
Password:
Save password
Enter
Forgotten password?
Register


Izvestiya: Mathematics, 2023, Volume 87, Issue 4, Pages 817–834
DOI: https://doi.org/10.4213/im9354e
(Mi im9354)
 

This article is cited in 3 scientific papers (total in 3 papers)

On stabilization of solutions of second-order semilinear parabolic equations on closed manifolds

D. V. Tunitsky

V. A. Trapeznikov Institute of Control Sciences of Russian Academy of Sciences, Moscow
References:
Abstract: The paper is concerned with problems of existence, uniqueness, and stabilization of weak solutions of one class of semilinear second-order parabolic differential equations on closed manifolds. These equations are inhomogeneous analogues of the Kolmogorov–Petrovskii–Piskunov–Fisher equation, and have significant applied and mathematical value.
Keywords: the Kolmogorov–Petrovskii–Piskunov–Fisher equation, second-order parabolic equation, semilinear equation on manifold, weak solution, stabilization.
Funding agency Grant number
Russian Science Foundation 19-11-00223
Russian Foundation for Basic Research 20-01-00610
Theorems 1, 3, 7 and 8 were proved with support of the Russian Science Foundation (project no. 19-11-00223); Theorems t2, 4, 5 and 6 were proved with support of the Russian Foundation for Basic Research (grant no. 20-01-00610a).
Received: 11.06.2022
Russian version:
Izvestiya Rossiiskoi Akademii Nauk. Seriya Matematicheskaya, 2023, Volume 87, Issue 4, Pages 186–204
DOI: https://doi.org/10.4213/im9354
Bibliographic databases:
Document Type: Article
UDC: 517.956.4+517.956.8+517.955
MSC: Primary 35L70; Secondary 35L60, 58A17
Language: English
Original paper language: Russian

Introduction

Semilinear second-order parabolic equations of the form

$$ \begin{equation} \frac{\partial q}{\partial t}+Lq=f(x,q), \end{equation} \tag{0.1} $$
where
$$ \begin{equation} Lq=-\sum_{l,m=1}^n\frac{\partial}{\partial x^l} \biggl(a^{l,m}(x)\, \frac{\partial q}{\partial x^m}\biggr)+ \sum_{l=1}^n b^l(x)\, \frac{\partial q}{\partial x^l} \end{equation} \tag{0.2} $$
is a linear elliptic differential operator, proved widely useful in mathematical modeling of various reaction–diffusion processes. Among numerous studies on these equations, we mention, first of all, the famous papers by Kolmogorov, Petrovskii, Piskunov [1] and Fisher [2], which dealt with a homogeneous operator $L$ and a homogeneous right-hand side $f$,
$$ \begin{equation*} L=-\sum_{l=1}^n\frac{\partial^2}{(\partial x^l)^2},\qquad f(x,q)=f(q). \end{equation*} \notag $$
We also mention the paper [3], which provides a historical account and a bibliographical survey on the studies of equations (0.1), (0.2). Some applications of these equations can be found in the book [4].

Equations (0.1), (0.2) in which the operator $L$ has periodic coefficients were studied in [3]. This case, which is reduced to an equation on a manifold diffeomorphic to the $n$-dimensional torus $\mathbb{T}^n$, has great applied value. Of special importance here are analogues of equations (0.1), (0.2) on other closed manifolds, and, in particular, on those diffeomorphic the $n$-dimensional sphere $\mathbb{S}^n$. In this regard, it is worth pointing out that, according to the generalized Poincaré conjecture, the sphere $\mathbb{S}^n$ is homeomorphic to each closed manifold homotopically equivalent to $\mathbb{S}^n$. The greatest challenge in this conjecture is in the dimension $n = 3$. In this case, the proof proposed by G. Ya. Perelman is based on the study of Ricci flows on closed three-dimensional manifolds (see [5]–[7]). This very approach proved capable of justifying the Poincaré conjecture, thereby solving one of the “millennium problems”. With the help of the DeTurck trick (see [8]), the problem of finding Ricci flows can be essentially reduced to solution of the corresponding parabolic equations. This clearly demonstrates that the study of solutions of nonlinear parabolic equations on closed manifolds has great applied and mathematical value. In the present paper, we consider existence, uniqueness, and stabilization of solutions of analogues of equation (0.1) on arbitrary closed finite-dimensional manifolds.

It is worth pointing out that, in many applied problems (for example, in many control problems), the right-hand sides of equations (0.1), (0.2) may involve terms which are not smooth or even not continuous. So, it is desirable to be able to select a class of admissible solutions based on which a satisfactory theory of solvability of such equations can be constructed under minimal requirements on the regularity of their coefficients. In the present paper, the weak solutions are considered as such a class. In this class, it proves possible to investigate solvability of analogues of the inhomogeneous equation (0.1) on closed manifolds under fairly light requirements on the regularity of their coefficients. In particular, some coefficients of these equations can be generalized function.

§ 1. Statement of the problem

1.1. Function spaces of tensor fields

By $X$ we denote an $n$-dimensional smooth closed Riemannian manifold, that is, a connected Hausdorff compact manifold without boundary equipped with metric $g\colon TX\times TX \to \mathbb{R}$. The metrics induced by $g$ on tensor bundles $(TX)^{\otimes^m}\otimes (T^* X)^{\otimes^l}$, $m,l=0,1,2,\dots$, of this manifold will be denoted by the same letter $g$; further, by $(TX)^{\otimes^0}\otimes (T^* X)^{\otimes^0}$ we will denote the trivial bundle $X \times \mathbb{R}$ with $g(r,t)=rt$ for $r,t \in \mathbb{R}$. The metric $g$ induces on the manifold $X$ the measure $V=V_g$ in the local coordinates by $x^1,\dots,x^n$

$$ \begin{equation} dV=\sqrt{g}\,dx^1\cdots dx^n, \end{equation} \tag{1.1} $$
where $g=\det\bigl(g(\partial/\partial x^m,\partial/\partial x^l)\bigr)$, and the Levi-Civita connection with the covariant differentiation operator $\nabla=\nabla_g$ uniquely defined by this connection.

Given real-value functions $u$ and $v$ on $X$, we set

$$ \begin{equation*} \langle u,v\rangle=\int_X u(x)v(x)\,dV, \qquad \operatorname*{ess\,sup}_{x \in X} u(x)= \inf_{\substack{S\subseteq X\\ V(S)=0}}\, \sup_{x \in X\setminus S} u(x). \end{equation*} \notag $$
From the metric $g$ and the measure $V$, one defines, in the usual way, the function space $L^p(X)$ and the space of tensor fields $L^p\bigl((TX)^{\otimes^m} \otimes (T^* X)^{\otimes^l}\bigr)$, where $p \geqslant 1$ and $m,l=0,1,2,\dots$ . Similarly, with the help of the covariant differentiation operator $\nabla$, one defines the Sobolev spaces $W^{k,p}(X)$ and $W^{k,p}\bigl((TX)^{\otimes^m}\otimes (T^* X)^{\otimes^l}\bigr)$, $k=0,1,2,\dots$, and the Hölder spaces $C^{k,\alpha}(X)$ and $C^{k,\alpha}\bigl((TX)^{\otimes^m}\otimes(T^* X)^{\otimes^l}\bigr)$ for $0<\alpha \leqslant 1$ (see [9], § 10.2.4, and [10], § 1).

On the tangent bundle of the Cartesian product $[0,T)\times X$, where $T \in (0,+\infty]$, consider the metric

$$ \begin{equation*} g_{[0,T)\times X}\colon T([0,T)\times X)\ni (\tau,\xi)\mapsto \tau^2+g(\xi,\xi)\in \mathbb{R} \end{equation*} \notag $$
and the corresponding measure $V_{g_{[0,T)\times X}}$ and the covariant differentiation operator $\nabla_{g_{[0,T)\times X}}$, using which, we similarly construct the function spaces
$$ \begin{equation*} \begin{gathered} \, L^p([0,T)\times X),\qquad L^p\bigl(\bigl(T([0,T) \times X)\bigr)^{\otimes^m} \otimes \bigl(T^*([0,T)\times X)\bigr)^{\otimes^l}\bigr), \\ W^{k,p}([0,T)\times X),\qquad W^{k,p}\bigl(\bigl(T([0,T) \times X)\bigr)^{\otimes^m} \otimes \bigl(T^*([0,T)\times X)\bigr)^{\otimes^l}\bigr), \\ C^{k,\alpha}([0,T)\times X),\qquad C^{k,\alpha} \bigl(\bigl(T([0,T) \times X)\bigr)^{\otimes^m} \otimes \bigl(T^*([0,T)\times X)\bigr)^{\otimes^l}\bigr). \end{gathered} \end{equation*} \notag $$

If $B$ is a Banach space with norm ${\|\,{\cdot}\,\|_B}$, then, for $T \in (0,+\infty]$, one defines in the standard way, the Banach spaces $L^p(T;B)$ with the norms

$$ \begin{equation*} \|q\|_{L^p(T;B)}=\biggl(\int_0^T\|q(t)\|_B^p\,dt\biggr)^{1/p}, \quad p\geqslant 1, \qquad \|q\|_{L^\infty(T;B)}= \operatorname*{ess\,sup}_{t\in [0,T)}\|q(t)\|_B, \end{equation*} \notag $$
see Chap. III, § 1 in [11] and Chap. II, § 2 in [12]. The spaces $L^2(T;W^{1,2}(X))$ and $W^{1,2}(T;L^2(X))$ are Hilbert spaces equipped with the inner products
$$ \begin{equation*} \begin{aligned} \, \langle q,p\rangle_{L^2(T;W^{1,2}(X))}&= \int_0^T\langle q(t),p(t)\rangle_{W^{1,2}(X)}\,dt, \\ \langle q,p\rangle_{W^{1,2}(T;L^2(X))}&= \int_0^T\bigl(\langle q(t),p(t)\rangle+ \langle q'(t),p'(t)\rangle\bigr)\,dt; \end{aligned} \end{equation*} \notag $$
their intersection $L^2(T;W^{1,2}(X))\cap W^{1,2}(T;L^2(X))$ is isomorphic to the Hilbert space $W^{1,2}([0,T)\times X)$ with the inner product
$$ \begin{equation*} \langle q,p\rangle_{W^{1,2}([0,T)\times X)}= \int_0^T\bigl(\langle q(t),p(t)\rangle+\langle dq(t),dp(t)\rangle_{L^2(T^* X)} +\langle q'(t),p'(t)\rangle\bigr)\,dt. \end{equation*} \notag $$
Here and in what follows, $q(t,\,\cdot\,)$ is identified with $q(t)$. Setting
$$ \begin{equation*} W(T;X)=L^2(T;W^{1,2}(X))\cap L^\infty(T;L^2(X)), \end{equation*} \notag $$
we get a Banach space with the norm
$$ \begin{equation*} \|q\|_{W(T;X)}^2=\operatorname*{ess\,sup}_{t\in [0,T)} \langle q(t),q(t)\rangle+\int_0^T\langle dq(t),dq(t)\rangle_{L^2(T^* X)}\,dt. \end{equation*} \notag $$

As usual, for a linear semilocal subspace $E \subseteq \mathcal{D}'([0,T)\times X)$, we let $E_{\mathrm{loc}}$ denote the smallest local subspace $\mathcal{D}'([0,T)\times X)$ containing $E$ (see § 10.1 in [13]). In particular, for $L^2([0,T)\times X)$ we have $L_{\mathrm{loc}}^2([0,T)\times X)$, for $L^\infty([0,T)\times X)$ we have $L_{\mathrm{loc}}^\infty([0,T)\times X)$, and for $W(T;X)$ we have $W_{\mathrm{loc}}(T;X)$.

1.2. Elliptic equations

In what follows, a particular property is said to hold almost everywhere if the set of points for which the property fails to hold is a set of measure zero with respect to the measure $V$ (see (1.1)). Let the Riemannian manifold $X$ be equipped, in addition to $g$, with another metric $a$. Assume that this metric is measurable, and there exist positive numbers $a_0$ and $a_1$ such that, almost everywhere,

$$ \begin{equation} a_0 g(\eta,\eta) \leqslant a(\eta,\eta) \leqslant a_1 g(\eta,\eta) \end{equation} \tag{1.2} $$
for all $\eta \in T^* X$. Consider the operators $d_a^*$ and $d_g^*$ formally adjoint with the exterior differentiation operator $d$ with respect to the metrics $a$ and $g$, respectively (see [14], Chap. VIII, § 1). In particular, $\langle a(du,v),1\rangle=\langle a(u,d_a^* v),1\rangle$ for all differential $k$-forms $u$ and $(k+1)$-forms $v$, $k=0,1,\dots,n-1$, and if the manifold $X$ is orientable, then, $d_a^*=(-1)^{n(k+1)+1}*d\,*$ on $k$-forms, where $* = *_a$ is the Hodge operator induced by the metric $a$. Given a function $u\in C^\infty(X)$, we define the linear second-order differential operator
$$ \begin{equation} Lu=\Delta u+bu+d_g^*(uc), \end{equation} \tag{1.3} $$
where $\Delta=\Delta_a=d_a^*\circ d$ is the geometric Laplacian (the Laplace–de Rahm operator; see Chap. IV, § 5 in [15]), and $b$, $c$ are measurable bounded (with respect to the metric $g$) vector field and linear differential form on $X$, respectively. In the local coordinates $x^1,\dots,x^n$,
$$ \begin{equation*} \begin{aligned} \, Lu&=-\frac{1}{\sqrt{a}} \sum_{l,m=1}^n \frac{\partial}{\partial x^l} \biggl(\sqrt{a}\,a(dx^l,dx^m)\, \frac{\partial u}{\partial x^m}\biggr)+ \sum_{l=1}^n b(dx^l)\,\frac{\partial u}{\partial x^l} \\ &\qquad-\frac{1}{\sqrt{g}} \sum_{l,m=1}^n \frac{\partial}{\partial x^l} \biggl(\sqrt{g}\,g(dx^l,dx^m) uc\biggl(\frac{\partial}{\partial x^m}\biggr)\biggr), \end{aligned} \end{equation*} \notag $$
where $a=\det\bigl(a(\partial/\partial x^m,\partial/\partial x^l)\bigr)$ and $g=\det\bigl(g(\partial/\partial x^m,\partial/\partial x^l)\bigr)$; cf. (0.2). In this context, condition (1.2) means that operator (1.3) is uniformly elliptic on the manifold $X$.

Consider the function

$$ \begin{equation} f\colon X \times \mathbb{R} \ni (x,r)\mapsto f(x,r)\in \mathbb{R} \end{equation} \tag{1.4} $$
which is locally Lipschitz continuous almost everywhere in $r$, that is, for any $r\in \mathbb{R}$ there exists a positive constant $\mu_0=\mu_0(r)$ such that, for $r_1,r_2 \in [-r,r]$,
$$ \begin{equation} |f(\,\cdot\, ,r_1)-f(\,\cdot\, ,r_2)| \leqslant \mu_0(r)|r_1-r_2| \end{equation} \tag{1.5} $$
almost everywhere. A weak (or generalized) solution of the equation
$$ \begin{equation} Lu=f+d_g^* h, \end{equation} \tag{1.6} $$
where $h$ is a measurable bounded linear differential form on the manifold $X$, is a function $u \in W^{1,2}(X)$ such that $f(\,\cdot\, ,u) \in L^2(X$), and
$$ \begin{equation} \mathcal{L}(u,v)=\langle f(\,\cdot\, ,u),v\rangle+ \langle h,dv\rangle_{L^2(T^* X)} \end{equation} \tag{1.7} $$
for each function $v \in C^\infty(X)$, where $\mathcal{L}$ is a continuous bilinear form
$$ \begin{equation} \mathcal{L}\colon W^{1,2}(X) \times C^\infty(X) \ni (u,v)\mapsto \langle a(du,dv),1\rangle+\langle bu,v\rangle+ \langle uc,dv\rangle_{L^2(T^* X)} \in \mathbb{R}. \end{equation} \tag{1.8} $$

Remark 1. According to (1.8), $\mathcal{L}(r,v) = r\langle c,dv\rangle_{L^2(T^* X)}$ for each $r \in \mathbb{R}$, and hence by definition (1.7) a constant function $u=r$ is a weak solution of equation (1.6) if and only if $\theta(r,v)=0$ for any function $v \in C^\infty(X)$, where

$$ \begin{equation} \theta \colon \mathbb{R} \times C^\infty(X) \ni (r,v)\mapsto r\langle c,dv\rangle_{L^2(T^* X)}-\langle f(\,\cdot\, ,r),v\rangle- \langle h,dv\rangle_{L^2(T^* X)} \in \mathbb{R}. \end{equation} \tag{1.9} $$
It is clear that any value of the functional $\theta$ is the difference between the left- and right-hand sides of equation (1.7) for $u=r$.

Remark 2. For a fixed $r$, the functional $\theta(r,\,\cdot\,)$ is a linear form on $C^\infty(X)$. The space $C^\infty(X)$ is dense in $W^{1,2}(X)$, and hence both the linear form $\theta(r,\,\cdot\,)$ (see (1.9)) and the bilinear form $\mathcal{L}$ (see (1.8)) can be uniquely extended by continuity to functions $v \in W^{1,2}(X)$. Moreover, if either of $\theta(r,v)=0$ or (1.7) holds for any function $v \in C^\infty(X)$, then it also holds for all functions $v \in W^{1,2}(X)$. It is clear that the fulfillment of any of these equalities for all $v \in W^{1,2}(X)$ is equivalent to that for only non-negative $v \in W^{1,2}(X)$.

1.3. Parabolic equations

The operator $L$ (1.3) is elliptic, and hence the evolutionary equation

$$ \begin{equation} \frac{\partial q}{\partial t}+Lq=f+d_g^* h \end{equation} \tag{1.10} $$
is parabolic. By a weak (or generalized) solution of equation (1.10) on the half-open interval $[0,T)$, $T \in (0,+\infty]$, with initial value
$$ \begin{equation} q(0)=q_0,\quad q_0\in L^2(X), \end{equation} \tag{1.11} $$
we mean a function $q \in W_{\mathrm{loc}}(T;X)$ such that $f(\,\cdot\, ,q) \in L_{\mathrm{loc}}^2([0,T) \times X)$, and
$$ \begin{equation} \langle q(t),p(t)\rangle+\mathcal{L}^t(q,p)=\langle q_0,p(0)\rangle+ \int_0^t\bigl(\langle f(\,\cdot\, ,q(\tau)),p(\tau)\rangle+ \langle h,dp(\tau)\rangle_{L^2 (T^* X)}\bigr)\,d\tau \end{equation} \tag{1.12} $$
for each $p \in C^\infty([0,T) \times X)$ and $t \in [0,T)$, where
$$ \begin{equation} \mathcal{L}^t \colon W(T;X) \times C^\infty([0,T)\times X) \ni (q,p) \mapsto \int_0^t\bigl(\mathcal{L}(q(\tau),p(\tau))- \langle q(\tau),p'(\tau)\rangle\bigr)\,d\tau \in \mathbb{R}. \end{equation} \tag{1.13} $$

Remark 3. By definitions (1.7) and (1.12), if the initial value $q_0$ (see (1.11)) is a solution of equation (1.6), then $q=q_0$ is a solution of the Cauchy problem (1.10), (1.11) on the infinite interval $[0,+\infty)$. Hence, in the case of a constant initial value $q_0=r \in \mathbb{R}$, according to Remark 1, the constant function $q=r$ is a weak solution of the Cauchy problem (1.10), (1.11) if and only if $\theta(r,v)=0$ for any function $v \in C^\infty(X)$.

Remark 4. The space $C^\infty([0,T) \times X)$ is dense in $W^{1,2}([0,T) \times X)$, and hence, as in Remark 2, the bilinear form $\mathcal{L}^t$ (1.13) can be uniquely extended by continuity to functions $p \in W^{1,2}([0,T) \times X)$ with traces $p(0),p(t) \in L^2(X)$. In addition, the fulfillment of (1.12) for $p \in W^{1,2}([0,T) \times X)$ is equivalent to that for only non-negative $p \in W^{1,2}([0,T) \times X)$.

§ 2. Main results

2.1. Existence, uniqueness, and regularity

Let us study conditions under which a weak solution of the Cauchy problem (1.10), (1.11) exists and is unique.

A function $u \in W^{1,2}(X)$ is a weak (or generalized) subsolution (supersolution) of equation (1.6) if $f(\,\cdot\, ,u) \in L^2(X)$ and if

$$ \begin{equation*} \mathcal{L}(u,v) \leqslant \langle f(\,\cdot\, ,u),v\rangle+ \langle h,dv\rangle_{L^2(T^* X)}\qquad \bigl(\mathcal{L}(u,v)\geqslant \langle f(\,\cdot\, ,u),v\rangle+ \langle h,dv\rangle_{L^2(T^* X)}\bigr) \end{equation*} \notag $$
for any non-negative function $v \in C^\infty(X)$. A function $q \in W_{\mathrm{loc}}(T;X)$ is a weak (or generalized) subsolution (supersolution) of the Cauchy problem (1.10), (1.11) on the half-open interval $[0,T)$, $T \in (0,+\infty]$, if $f(\,\cdot\, ,q)\in L_{\mathrm{loc}}^2([0,T) \times X)$ and, if, for any non-negative function $p \in C^\infty([0,T) \times X)$,
$$ \begin{equation} \begin{aligned} \, &\langle q(t),p(t)\rangle+\mathcal{L}^t(q,p) \\ &\quad\leqslant \langle q_0,p(0)\rangle \,{+}\int_0^t\bigl(\langle f(\,\cdot\, ,q(\tau)),p(\tau)\rangle \,{+}\,\langle h,dp(\tau)\rangle_{L^2(T^* X)}\bigr)\,d\tau \\ &\biggl(\langle q(t),p(t)\rangle+ \mathcal{L}^t(q,p) \\ &\quad\geqslant\langle q_0,p(0)\rangle \,{+}\int_0^t\bigl(\langle f(\,\cdot\, ,q(\tau)),p(\tau)\rangle \,{+}\,\langle h,dp(\tau)\rangle_{L^2(T^* X)}\bigr)\,d\tau\biggr),\qquad t \in [0,T). \end{aligned} \end{equation} \tag{2.1} $$
A weak subsolution (supersolution) of problem (1.10), (1.11) which is not a weak solution will be called strong. In what follows, unless otherwise explicitly stipulated, we will assume that all solutions, subsolutions, and supersolutions are weak, and the adjective “weak” will be dropped for brevity.

Remark 5. It is clear that each solution is simultaneously a subsolution and a supersolution. Conversely, if a function is simultaneously a subsolution and a supersolution, then by the concluding lines of Remarks 2 and 4, this function is a solution. It is easily checked that if $w \in W^{1,2}(X)$ is a subsolution (supersolution) of equation (1.6) and $w \leqslant q_0$ ($w \geqslant q_0$) almost everywhere, then $q=w$ is a subsolution (supersolution) of problem (1.10), (1.11) on the infinite interval $[0,+\infty)$. Next, in analogy with Remark 1, the constant function $u=r$, $r \in \mathbb{R}$, is a subsolution (supersolution) of equation (1.6) if and only if $\theta(r,v) \leqslant 0$ ($\theta(r,v) \geqslant 0$) for each non-negative function $v \in C^\infty(X)$. Correspondingly, in analogy with Remark 3, if $r \leqslant q_0$ ($r \geqslant q_0$) almost everywhere and $\theta(r,v) \leqslant 0$ ($\theta(r,v) \geqslant 0$) for each non-negative function $v \in C^\infty(X)$, then $q=r$ is a subsolution (supersolution) of problem (1.10), (1.11).

The following result holds.

Theorem 1 (existence and uniqueness of a solution). Assume that a metric $a\in L^\infty\bigl((T^* X)^{\otimes^2}\bigr)$ satisfies estimate (1.2), a vector field $b$ lies in $L^\infty(TX)$, differential forms $c,h $ lie in $L^\infty(T^* X)$, there exists a number $\mu \in \mathbb{R}$ such that

$$ \begin{equation} \langle c,dv\rangle_{L^2(T^* X)}+\langle \mu,v\rangle \geqslant 0 \end{equation} \tag{2.2} $$
for any non-negative function $v \in C^\infty(X)$, and $f \in L_{\mathrm{loc}}^\infty(X \times \mathbb{R})$ satisfies the Lipschitz condition (see (1.5)) almost everywhere. Let $q_1$ and $q_2$ be, respectively, a subsolution and supersolution of problem (1.10), (1.11) on $[0,T)$, $T \in (0,+\infty]$, $q_1,q_2 \in L^\infty([0,T) \times X)$. Then on $[0,T)$ there exists a unique solution $q$ of this problem, and $q_1(t) \leqslant q(t) \leqslant q_2(t)$ almost everywhere for $t \in [0,T)$.

For a proof, see § 3.

It is known that solutions of the Cauchy problem (1.10), (1.11) are locally Hölder-continuous. More precisely, the following result holds.

Theorem 2 (regularity of the solution). Let the coefficients and right-hand side of (1.10) satisfy all the conditions of Theorem 1 except (2.2). Let $q$ be a solution of the Cauchy problem (1.10), (1.11) from $L^\infty([0,T) \times X)$, $T \in (0,+\infty)$. Then $q \in C(T;L^2(X))$ and, for each $(t,x) \in (0,T] \times X$, there exist its neighbourhood $U$ and a number $0<\alpha<1$ such that $q_{|U\cap [0,T) \times X}\in C^{0,\alpha}(U \cap [0,T)\times X)$.

Under the above constraints on the coefficients of the differential operator $L$ (see (1.3)) and the function $f$ (see (1.4)), the result of Theorem 2 follows from the known properties of solutions of linear parabolic equations (see Chap. VI, § 7 in [16] and § 1.5 in [17]). It is natural that a further increase of regularity of the coefficients of equation (1.10) increases correspondingly the regularity of its solutions (see Chap. VI, § 2 in [16]).

The following result is clear from Theorems 1 and 2.

Theorem 3 (global existence and uniqueness). Let the coefficients and right-hand side of equation (1.10) satisfy the conditions of Theorem 1. If $q_1$ and $q_2$ are, respectively, a subsolution and a supersolution of problem (1.10), (1.11) on $[0,T)$, $T \in (0,+\infty]$, and $q_1,q_2 \in L_{\mathrm{loc}}^\infty ([0,T) \times X)$, then on $[0,T)$ problem (1.10), (1.11) has a unique solution $q$, which is locally Hölder-continuous on $(0,T) \times X$, $q \in C(T;L^2(X))$, and $q_1(t) \leqslant q(t) \leqslant q_2(t)$ almost everywhere for $t \in [0,T)$.

2.2. Stabilization of solutions

Let us find conditions under which the solution $q$ of the Cauchy problem (1.10), (1.11) tends as $t\to +\infty$ to the solution $u$ of the stationary equation (1.6).

A function $f$ (see (1.4)) is called strictly right (left) concave at a point $r_0 \in \mathbb{R}$ if

$$ \begin{equation} (1-\alpha)f(\,\cdot\, ,r_0)+\alpha f(\,\cdot\, ,r)< f\bigl(\,\cdot\, ,(1-\alpha) r_0+\alpha r\bigr) \end{equation} \tag{2.3} $$
almost everywhere for $r>r_0$ ($r<r_0$) and $0<\alpha<1$. Similarly, a function $f$ is called strictly right (left) convex at a point $r_0 \in \mathbb{R}$ if
$$ \begin{equation} f\bigl(\,\cdot\, ,(1-\alpha)r_0+\alpha r\bigr)<(1-\alpha)f(\,\cdot\, ,r_0)+ \alpha f(\,\cdot\, ,r) \end{equation} \tag{2.4} $$
almost everywhere for $r>r_0$ ($r<r_0$) and $0<\alpha<1$. The solutions of problem (1.10), (1.11) have the following asymptotic properties.

Theorem 4 (stabilization to a stationary solution). Let the coefficients and right-hand side of equation (1.10) satisfy the conditions of Theorem 1, let $r_0,N \in \mathbb{R}$, $r_0 \ne N$, and let functions $q_0 \in L^2(X)$ and $w \in W^{1,2}(X)\cap L^\infty(X)$ be such that $w \geqslant 0$ almost everywhere, $V(x \in X\mid w \ne 0)>0$ and $V(x \in X \mid q_0 \ne r_0)>0$.

Then the following hold.

(a) Let $f$ be strictly right concave at $r_0$, let $r_0$ and $N$ be, respectively, a subsolution and supersolution of equation (1.6), let the initial value (1.11) satisfy $r_0 \leqslant q_0 \leqslant N$ almost everywhere, and let there exist $\delta$ such that $r_0+\alpha w$ is a subsolution of equation (1.6) for $0<\alpha<\delta$. Then both equation (1.6) and the Cauchy problem (1.10), (1.11) have unique solutions $u$ and $q$, which are continuous, $r_0< u, q(t) \leqslant N$ for $0<t<+\infty$, and satisfy the limit relation

$$ \begin{equation} \lim_{t \to \infty}\|q(t)-u\|_{C(X)}=0. \end{equation} \tag{2.5} $$

(b) Let $f$ be strictly left convex at $r_0$, let $N$ and $r_0$ be, respectively, a subsolution and supersolution of equation (1.6), let the initial value (1.11) satisfy $N \leqslant q_0 \leqslant r_0$ almost everywhere, and let there exist $\delta$ such that $r_0-\alpha w$ is a supersolution of equation (1.6) for $0<\alpha<\delta$. Then both equation (1.6) and the Cauchy problem (1.10), (1.11) have unique solutions $u$ and $q$, which are continuous, $N \leqslant u,q(t)<r_0$ for $0<t<+\infty$, and satisfy (2.5).

For a proof, see § 4.

Remark 6. The idea of the proof is to construct a non-increasing (in $t$) function $\widetilde{q} \in C([0,+\infty)\times X)$ such that

$$ \begin{equation} |q(t)-u| \leqslant \widetilde{q}(t), \qquad \lim_{t\to\infty}\|\widetilde{q}(t)\|_{C(X)}=0. \end{equation} \tag{2.6} $$

Theorem 5 (stabilization to a constant solution). Let the coefficients and right-hand side of equation (1.10) satisfy the conditions of Theorem 1, let $r_0,A \in \mathbb{R}$, $A>0$, let $q_0 \in L^2(X)$, $w \in W^{1,2}(X) \cap L^\infty(X)$, and let $w \geqslant 0$ almost everywhere

Then the following hold.

(a) Let the initial value (1.11) satisfy $r_0 \leqslant q_0 \leqslant r_0+Aw$ almost everywhere, let $r_0$ be a solution and $r_0+\alpha w$ be a strong supersolution of equation (1.6) for $0<\alpha \leqslant A$. Then the Cauchy problem (1.10), (1.11) has a unique solution $q$, which is continuous, $r_0 \leqslant q \leqslant r_0+Aw$ almost everywhere for $0<t<+\infty$, and a unique solution $u$ of equation (1.6) satisfying $r_0 \leqslant u \leqslant r_0+Aw$ almost everywhere is $u=r_0$, and the limit relation holds

$$ \begin{equation} \lim_{t\to +\infty}\|q(t)-r_0\|_{C(X)}=0. \end{equation} \tag{2.7} $$

(b) Let the initial value (1.11) satisfy $r_0-Aw \leqslant q_0 \leqslant r_0$ almost everywhere, let $r_0$ be a solution and $r_0-\alpha w$ be a strong subsolution of equation (1.6) for $0<\alpha \leqslant A$. Then the Cauchy problem (1.10), (1.11) has a unique solution $q$, which is continuous, $r_0-Aw \leqslant q \leqslant r_0$ almost everywhere for $0<t<+\infty$, and a unique solution $u$ of equation (1.6) satisfying $r_0-Aw \leqslant u \leqslant r_0$ almost everywhere is $u=r_0$, and the limit relation (2.7) holds.

For a proof, see § 4; moreover, Remark 6 with $u=r_0$ applies as in Theorem 4.

Remark 7. Under the conditions of Theorems 4 and 5, using standard a priori estimates of solutions of linear second-order parabolic equations (see Chap. VI, § 1 in [16] and § 1.5 in [17]) and applying Theorems 4 and 5, it can be shown that the solutions of problem (1.10), (1.11) are asymptotically stable.

2.3. The eigenvalue problem

Assume that the function $f$ (see (1.4)) has the right (left) derivative

$$ \begin{equation*} \begin{gathered} \, \frac{\partial f}{\partial r}(\,\cdot\, ,r_0+0)=\lim_{r \to r_0+0} \frac{f(\,\cdot\, ,r)-f(\,\cdot\, ,r_0)}{r-r_0} \\ \biggl(\frac{\partial f}{\partial r}(\,\cdot\, ,r_0-0)= \lim_{r\to r_0-0}\frac{f(\,\cdot\, ,r)-f(\,\cdot\, ,r_0)}{r-r_0}\biggr), \end{gathered} \end{equation*} \notag $$
almost everywhere at a point $r_0 \in \mathbb{R}$. For the differential operator $L$ (see (1.3)), consider the eigenvalue problem
$$ \begin{equation} Lw-Fw=\lambda w \end{equation} \tag{2.8} $$
with the potential
$$ \begin{equation} F(x)=\frac{\partial f}{\partial r}(x,r_0+0)\qquad \biggl(F(x)=\frac{\partial f}{\partial r}(x,r_0-0)\biggr). \end{equation} \tag{2.9} $$
We will say that $w \in W^{1,2}(X)$ is an eigenfunction belonging to an eigenvalue $\lambda \in \mathbb{R}$ if it satisfies equality (2.8) in the weak sense, that is,
$$ \begin{equation} \langle Lw-Fw,v\rangle=\lambda\langle w,v\rangle \end{equation} \tag{2.10} $$
for all functions $v \in C^\infty(X)$. The following result is known.

Theorem 6. Let the coefficients of the operator $L$ (see (1.3)) satisfy the conditions of Theorem 1. If $F \in L^\infty(X)$, then problem (2.8)(2.10) has a unique simple eigenvalue $\lambda_1$ to which there belongs a real eigenfunction $w \in C^{0,\alpha}(X)$, $0<\alpha<1$, which does not vanish on $X$.

The required result follows from the Krein–Rutman theorem (see § 6.5.2 in [18], § 6.4 in [19], Chap. 11, § C in [20], and Chap. I, § 7 in [21]) and from the regularity of solutions of elliptic equations (see § 8.9 in [22]).

The eigenvalue $\lambda_1$ is known as the principal eigenvalue.

From definition (2.3) it easily follows that the function $f$ (see (1.4)) is strictly right (left) concave at a point $r_0 \in \mathbb{R}$ if and only if

$$ \begin{equation} \frac{f(\,\cdot\, ,r_1)-f(\,\cdot\, ,r_0)}{r_1-r_0}< \frac{f(\,\cdot\, ,r)-f(\,\cdot\, ,r_0)}{r-r_0} \end{equation} \tag{2.11} $$
almost everywhere for $r_0<r<r_1$ ($r_1<r<r_0$). Similarly, from definition (2.4) it easily follows that a function $f$ is strictly right (left) convex at a point $r_0 \in \mathbb{R}$ if and only if
$$ \begin{equation} \frac{f(\,\cdot\, ,r_1)-f(\,\cdot\, ,r_0)}{r_1-r_0}> \frac{f(\,\cdot\, ,r)-f(\,\cdot\, ,r_0)}{r-r_0} \end{equation} \tag{2.12} $$
almost everywhere for $r_0<r<r_1$ ($r_1<r<r_0$). The monotonicity properties of the difference relations (2.11) and (2.12) imply that the derivatives in the right-hand sides of (2.9) exist and lie in $L^\infty(X)$. As a corollary, problem (2.8)(2.10) is well posed.

The following results hold.

Theorem 7. Let the coefficients and right-hand side of equation (1.10) satisfy the conditions of Theorem 1, and let $\lambda_1<0$ be the principal eigenvalue. Then the following hold.

(a) Let $f$ be strictly right concave at $r_0 \in \mathbb{R}$, let $r_0$ and $N$, $r_0<N$, be, respectively, a subsolution and supersolution of equation (1.6), and let the initial value (1.11) satisfy $r_0\leqslant q_0\leqslant N$ almost everywhere and $V(x\in X\mid q_0 \ne r_0)>0$. Then both equation (1.6) and the Cauchy problem (1.10), (1.11) have unique solutions $u$ and $q$, which are continuous, $r_0<u,q(t)\leqslant N$ for $0<t<+\infty$, and satisfy the limit relation (2.5).

(b) Let $f$ be strictly left convex at $r_0 \in \mathbb{R}$, let $r_0$ and $N$, $N<r_0$, be, respectively, a supersolution and subsolution of equation (1.6), and let the initial value (1.11) satisfy $N\leqslant q_0\leqslant r_0$ almost everywhere and $V(x\in X\mid q_0 \ne r_0)>0$. Then both equation (1.6) and the Cauchy problem (1.10), (1.11) have unique solutions $u$ and $q$, which are continuous, $N \leqslant u,q(t)<r_0$ for $0<t\leqslant +\infty$, and satisfy the limit relation (2.5).

For a proof, see § 5; moreover, Remark 6 applies as in Theorem 4.

Theorem 8. Let the coefficients and right-hand side of equation (1.10) satisfy the conditions of Theorem 1, let a number $r_0 \in \mathbb{R}$ be a solution of equation (1.6), and let $q_0 \in L^\infty(X)$ and $\lambda_1 \geqslant0$.

Then the following hold.

(a) Let $f$ be strictly right concave at $r_0$ and the initial value (1.11) satisfy the inequality $q_0 \geqslant r_0$ almost everywhere. Then the Cauchy problem (1.10), (1.11) has a unique solution $q$, which is continuous, $r_0 \leqslant q(t)$ for $0<t\leqslant +\infty$, and equation (1.6) has a unique solution $u$ such that $r_0 \leqslant u$ almost everywhere, $u=r_0$, and which satisfies (2.7).

(b) Let $f$ be strictly left concave at $r_0$ and let the initial value (1.11) satisfy $q_0 \leqslant r_0$ almost everywhere. Then the Cauchy problem (1.10), (1.11) has a unique solution $q$, which is continuous, satisfies $q(t)\leqslant r_0$ for $0<t\leqslant +\infty$, and equation (1.6) has a unique solution $u$ such that $u\leqslant r_0$ almost everywhere, $u=r_0$, and which satisfies (2.7).

For a proof, see § 5. In addition, Remark 6 with $u=r_0$ applies as in Theorem 4. Moreover, similarly to Remark 7, under the conditions of Theorems 7 and 8, the solutions of problem (1.10), (1.11) are asymptotically stable.

§ 3. Existence and uniqueness of solutions

3.1. Auxiliary results

For a proof of Theorem 1, we will require two technical results. The first one is the following comparison test.

Lemma 1. Let the coefficients of the differential operator $L$ (see (1.3)) and the function $f$ (see (1.4)) satisfy the conditions of Theorem 1. If $q_1,q_2\in W(T;X)\cap L^\infty([0,T) \times X)$, $T\in (0,+\infty]$, and non-negative function $p\in C^\infty([0,T) \times X)$ satisfy

$$ \begin{equation*} \begin{aligned} \, &\langle q_1(t),p(t)\rangle+\mathcal{L}^t(q_1,p)-\langle q_1(0),p(0)\rangle- \int_0^t\langle f(\,\cdot\, ,q_1(\tau)),p(\tau)\rangle \,d\tau \\ &\qquad\leqslant \langle q_2(t),p(t)\rangle+\mathcal{L}^t(q_2,p)- \langle q_2(0),p(0)\rangle- \int_0^t\langle f(\,\cdot\, ,q_2(\tau)),p(\tau)\rangle\,d\tau \end{aligned} \end{equation*} \notag $$
for $t\in[0,T)$, and $q_1 (0)\leqslant q_2 (0)$ almost everywhere, then $q_1 (t)\leqslant q_2 (t)$ almost everywhere for $t\in[0,T)$. In addition, if $V\bigl(\{x \in X\mid q_1(0) \ne q_2(0)\}\bigr)>0$, then $q_1(t)<q_2(t)$ almost everywhere for $t\in(0,T)$.

Proof. By the Lipschitz condition (1.5), if
$$ \begin{equation*} \mu_0=\mu_0(\max\{\|q_1\|_{L^\infty([0,T) \times X)}, \|q_2\|_{L^\infty([0,T) \times X)}\}) \end{equation*} \notag $$
then
$$ \begin{equation*} f(\,\cdot\, ,q_1(\tau))-f(\,\cdot\, ,q_2(\tau))\leqslant \mu_0\bigl(q_1(\tau)-q_2(\tau)\bigr) \operatorname{sgn}\bigl(q_1(\tau)-q_2(\tau)\bigr) \end{equation*} \notag $$
almost everywhere. Hence, by the assumption, for $q=q_1-q_2$ and for a non-negative $p\in C^\infty([0,T) \times X)$, we have
$$ \begin{equation*} \langle q(t),p(t)\rangle+\mathcal{L}^t(q,p)-\langle q(0),p(0)\rangle- \int_0^t\langle \mu_0 q(\tau)\operatorname{sgn}q(\tau), p(\tau)\rangle\,d\tau \leqslant 0. \end{equation*} \notag $$
By definition of $\mathcal{L}$ (see (1.8)), $\mathcal{L}(e^{\mu t}q,p)=\mathcal{L}(q,e^{\mu t}p)$. Hence, by definition of $\mathcal{L}^t$ (see (1.13)) and substituting $q=e^{\mu t}\widetilde{q}$ and $\widetilde{p}=e^{\mu t}p$, the last inequality assumes the form
$$ \begin{equation*} \langle \widetilde{q}(t),\widetilde{p}(t)\rangle+ \mathcal{L}^t(\widetilde{q},\widetilde{p})- \langle q(0),\widetilde{p}(0)\rangle+\int_0^t \langle(\mu-\mu_0\operatorname{sgn}q(\tau)) \widetilde{q}(\tau),\widetilde{p}(\tau)\rangle\,d\tau \leqslant0 \end{equation*} \notag $$
with non-negative $\widetilde{p}\in C^\infty([0,T) \times X)$. For sufficiently large $\mu>0$, the conclusion of the lemma follows from inequality (2.2) and the strong parabolic maximum principle (see Chap. VI, § 7 in [16] and § 1.5 in [17]). Lemma is proved.

Note that that conclusion of Lemma 1 is true, in particular, if $q_1(0)=q_2(0)$ almost everywhere, and also if $f \equiv 0$.

The second technical result we require is the following a priori estimate.

Lemma 2. Let the coefficients of the differential operator $L$ (see (1.3)) satisfy the conditions of Theorem 1. If $q_0 \in L^2(X)$, $q,v \in W(T;X)$, $T\in (0,+\infty]$, and $p\in C^\infty([0,T) \times X)$ satisfy

$$ \begin{equation*} \langle q(t),p(t)\rangle+\mathcal{L}^t(q,p)=\langle q_0,p(0)\rangle+ \int_0^t\langle v(\tau),p(\tau)\rangle\,d\tau,\qquad t\in[0,T), \end{equation*} \notag $$
then there exists a constant $C>0$, independent of $q_0$ and $v$, such that
$$ \begin{equation*} \|q\|_{W(t;X)} \leqslant Ce^{Ct}(\|v\|_{L^2([0,t) \times X)}+ \|q_0\|_{L^2(X)}),\qquad t\in [0,T). \end{equation*} \notag $$

A proof of Lemma 2 follows from the known a priori estimates for solutions of linear second-order parabolic equations (see Chap. VI, § 1 in [16] and § 1.5 in [17]).

Note that, in particular, Lemma 2 holds with $q_0 \equiv 0$.

3.2. Proof of Theorem 1

By the assumption of the theorem and definition (2.1),

$$ \begin{equation*} \begin{aligned} \, &\langle q_1(t),p(t)\rangle+\mathcal{L}^t(q_1,p)-\langle q_0,p(0)\rangle- \int_0^t\langle f(\,\cdot\, ,q_1(\tau)),p(\tau)\rangle\,d\tau \\ &\qquad\leqslant \langle q_2(t),p(t)\rangle+ \mathcal{L}^t(q_2,p)-\langle q_0,p(0)\rangle- \int_0^t\langle f(\,\cdot\, ,q_2(\tau)),p(\tau)\rangle\,d\tau \end{aligned} \end{equation*} \notag $$
for non-negative $p \in C^\infty([0,T) \times X)$ and $t \in [0,T)$. By Lemma 1, we have $q_1(t) \leqslant q_2(t)$ almost everywhere for $t\in[0,T)$. The function $f$ (see (1.4)) satisfies condition (1.5) almost everywhere. Hence, if
$$ \begin{equation*} \mu \geqslant \mu_0(\max\{\|q_1\|_{L^\infty([0,T) \times X)}, \|q_2\|_{L^\infty([0,T) \times X)}\}) \end{equation*} \notag $$
and if functions $v_1$ and $v_2$ are such that $q_1\leqslant v_1 \leqslant v_2 \leqslant q_2$ almost everywhere, then
$$ \begin{equation*} f(\,\cdot\, ,v_1)+\mu v_1 \leqslant f(\,\cdot\, ,v_2)+\mu v_2 \end{equation*} \notag $$
almost everywhere (cf. the appendix to Chap. IV, § 2 in [23] and [24]), and, therefore,
$$ \begin{equation} \langle f(\,\cdot\, ,v_1)+\mu v_1,v\rangle \leqslant \langle f(\,\cdot\, ,v_2)+\mu v_2,v\rangle \end{equation} \tag{3.1} $$
for any non-negative function $v \in C^\infty(X)$. It is clear that (1.10) is equivalent to the equation
$$ \begin{equation*} \frac{\partial q}{\partial t}+Lq+\mu q=f(\,\cdot\, ,q)+d_g^* h+\mu q. \end{equation*} \notag $$

Let us construct successive approximations $\{q_{1,l}\}$ and $\{q_{2,l}\}$ by induction. We set $q_{1,0}=q_1$, $q_{2,0}=q_2$, and as $q_{j,l}$ for $j=1,2$ and $l=1,2,\dots$ we take the solution of the Cauchy problem

$$ \begin{equation*} \frac{\partial q_{j,l}}{\partial t}+Lq_{j,l}+\mu q_{j,l}= f(\,\cdot\, ,q_{j,l-1})+d_g^* h+\mu q_{j,l-1},\qquad q_{j,l}(0)=q_0, \end{equation*} \notag $$
$l=1,2,\dots$, that is, by definition (1.12), for any function $p\in C^\infty([0,T) \times X)$ and $t\in [0,T)$, we have
$$ \begin{equation} \begin{aligned} \, &\langle q_{j,l}(t),p(t)\rangle+\mathcal{L}^t(q_{j,l},p)+ \mu\int_0^t\langle q_{j,l}(\tau),p(\tau)\rangle\,d\tau \nonumber \\ &\ =\langle q_0,p(0)\rangle+\int_0^t \bigl(\langle f(\,\cdot\, ,q_{j,l-1}(\tau))+\mu q_{j,l-1}(\tau),p(\tau)\rangle+ \langle h,dp(\tau)\rangle_{L^2(T^* X)}\bigr)\,d\tau. \end{aligned} \end{equation} \tag{3.2} $$
The linear equation (3.2) has a unique solution $q_{j,l}\in W(T;X)$, see Chap. VI, § 1 in [16] and § 1.5 in [17]. Hence, by monotonicity (3.1) and by construction (3.2),
$$ \begin{equation*} \begin{aligned} \, &\langle q_{j,l}(t),p(t)\rangle+\mathcal{L}^t(q_{j,l},p)+ \mu\int_0^t\langle q_{j,l}(\tau),p(\tau)\rangle\,d\tau \\ &\qquad\leqslant \langle q_{j,l-1}(t),p(t)\rangle+ \mathcal{L}^t(q_{j,l-1},p)+ \mu\int_0^t\langle q_{j,l-1}(\tau),p(\tau)\rangle\,d\tau. \end{aligned} \end{equation*} \notag $$
Hence by Lemma 1, for $t\in[0,T)$,
$$ \begin{equation} q_1(t)\leqslant q_{1,1}(t) \leqslant q_{1,2}(t)\leqslant\dots\leqslant q_{2,2}(t) \leqslant q_{2,1}(t) \leqslant q_2(t) \end{equation} \tag{3.3} $$
almost everywhere. By monotonicity and since $\{q_{1,l}\}$ and $\{q_{2,l}\}$ are bounded, the limits
$$ \begin{equation} q_{1,\infty}=\lim_{l \to +\infty}q_{1,l},\qquad q_{2,\infty}=\lim_{l\to +\infty}q_{2,l} \end{equation} \tag{3.4} $$
exist almost everywhere and satisfy
$$ \begin{equation} q_1(t) \leqslant q_{1,1}(t) \leqslant q_{1,2}(t)\leqslant\dots\leqslant q_{1,\infty}(t) \leqslant q_{2,\infty}(t)\leqslant\dots\leqslant q_{2,2}(t) \leqslant q_{2,1}(t) \leqslant q_2(t) \end{equation} \tag{3.5} $$
almost everywhere. Hence by the Lebesgue dominated convergence theorem, the limit relations (3.4) also hold in the norm $\|\,{\cdot}\,\|_{L^2([0,t)\times X)}$ for $t\in [0,T)$. Next, by construction (3.2)
$$ \begin{equation*} \begin{aligned} \, &\langle q_{j,l+m}(t)-q_{j,l}(t),p(t)\rangle+\mathcal{L}^t(q_{j,l+m}- q_{j,l},p)+\mu\int_0^t\langle q_{j,l+m}(\tau)-q_{j,l}(\tau),p(\tau)\rangle\,d\tau \\ &\quad =\int_0^t\langle f(\,\cdot\, ,q_{j,l-1+m}(\tau))- f(\,\cdot\, ,q_{j,l-1}(\tau))+\mu(q_{j,l-1+m}(\tau) -q_{j,l-1}(\tau)),p(\tau)\rangle\,d\tau, \end{aligned} \end{equation*} \notag $$
and hence by Lemma 2 there exists $C>0$ such that, for $j=1,2$ and $l,m=1,2,\dots$,
$$ \begin{equation*} \begin{aligned} \, \|q_{j,l+m}-q_{j,l}\|_{W(t;X)} &\leqslant Ce^{Ct}\|f(\,\cdot\, ,q_{j,l-1+m})- f(\,\cdot\, ,q_{j,l-1})\|_{L^2([0,t)\times X)} \\ &\qquad+\mu Ce^{Ct}(\|q_{j,l-1+m}-q_{j,l-1}-q_{j,l+m}+ q_{j,l}\|_{L^2([0,t) \times X)}). \end{aligned} \end{equation*} \notag $$
By the already established $\|\,{\cdot}\,\|_{L^2([0,t)\times X)}$-convergence of the approximations $\{q_{1,l}\}$ and $\{q_{2,l}\}$ and also from (3.3) and (1.5), it follows that $\{q_{1,l}\}$ and $\{q_{2,l}\}$ are Cauchy sequences also in the norm $\|\,{\cdot}\,\|_{W(t;X)}$ (see (1.2)). Since $W(t;X)$ is complete, these sequences converge to the functions $q_{1,\infty}$ and $q_{2,\infty}$ (see (3.4)). Therefore, by passing to a limit in (3.2), we get
$$ \begin{equation*} \begin{aligned} \, &\langle q_{j,\infty}(t),p(t)\rangle+L^t(q_{j,\infty},p)= \langle q_0,p(0)\rangle \\ &\qquad\qquad+\int_0^t\bigl(\langle f(\,\cdot\, ,q_{j,\infty}(\tau)),p(\tau)\rangle+ \langle h,dp(\tau)\rangle_{L^2(T^* X)}\bigr)\,d\tau \end{aligned} \end{equation*} \notag $$
for $j=0,1$, $p\in C^\infty([0,T) \times X)$ and $t\in[0,T)$, that is, $q_{j,\infty}$ are solutions of problem (1.10), (1.11), and, by (3.5), $q_1(t) \leqslant q_{1,\infty}(t) \leqslant q_{2,\infty}(t) \leqslant q_2(t)$ almost everywhere.

By Lemma 1, any solution $q$ of problem (1.10), (1.11) satisfies $q_{1,\infty}=q=q_{2,\infty}$ almost everywhere on the half-open interval $[0,T)$. Theorem 1 is proved.

§ 4. Stabilization of solutions

4.1. Auxiliary result

In the proof of Theorems 4 and 5 we will require the following auxiliary result.

Lemma 3. Let the coefficients and right-hand side of equation (1.10) satisfy the conditions of Theorem 1, and let $w_1$ and $w_2$ be, respectively, a subsolution and supersolution of equation (1.6) such that $w_1,w_2 \in L^\infty(X)$ and $w_1 \leqslant w_2$ almost everywhere. Then there exist unique solutions $q_1,q_2 $ of the problems

$$ \begin{equation} \frac{\partial q_1}{\partial t}+Lq_1 =f(\,\cdot\, ,q_1)+d_g^* h, \qquad q_1(0) =w_1, \end{equation} \tag{4.1} $$
$$ \begin{equation} \frac{\partial q_2}{\partial t}+Lq_2 =f(\,\cdot\, ,q_2)+d_g^* h, \qquad q_2(0) =w_2. \end{equation} \tag{4.2} $$
Moreover, the solutions $q_1,q_2 $ are locally Hölder-continuous on $(0,+\infty)\times X$, $q_1,q_2 \in C(+\infty;L^2(X))$, and $w_1 \leqslant q_1(t) \leqslant q_2(t) \leqslant w_2$ almost everywhere for $t\in[0,+\infty)$. Furthermore, $q_1$ is non-decreasing, and $q_2$ is non-increasing in $t$, and both $q_1,q_2$ converge uniformly to the continuous solutions $q_{1,\infty}$ and $q_{2,\infty}$ of equation (1.6):
$$ \begin{equation} q_{1,\infty} =\lim_{t\to +\infty}q_1(t), \end{equation} \tag{4.3} $$
$$ \begin{equation} q_{2,\infty} =\lim_{t\to +\infty}q_2(t). \end{equation} \tag{4.4} $$
So, the solutions $q_1$, $q_{1,\infty}$, $q_2$ and $q_{2,\infty}$ satisfy
$$ \begin{equation*} \lim_{t \to +\infty}\|q_1(t)-q_{1,\infty}\|_{C(X)}=0,\qquad \lim_{t \to +\infty}\|q_2(t)-q_{2,\infty}\|_{C(X)}=0, \end{equation*} \notag $$
and, for $0 \leqslant t' \leqslant t''$, the estimates
$$ \begin{equation} w_1 \leqslant q_1(t') \leqslant q_1(t'') \leqslant q_{1,\infty} \leqslant q_{2,\infty} \leqslant q_2(t'') \leqslant q_2(t') \leqslant w_2 \end{equation} \tag{4.5} $$
hold almost everywhere.

Proof. By the conditions of the lemma and in view of Remark 5, the functions $w_1$ and $w_2$ are, respectively, a subsolution and a supersolution of problem (1.10), (1.11). Therefore, by Theorem 3, on $[0,+\infty)$, there exist unique solutions $q_1$ and $q_2$ of the Cauchy problems (4.1) and (4.2). Moreover, $q_1,q_2 \in C(+\infty;L^2(X))$, $q_1,q_2$ are locally Hölder-continuous on $(0,+\infty)\times X$, and $w_1 \leqslant q_1(t) \leqslant q_2(t) \leqslant w_2$ almost everywhere for $t\in[0,+\infty)$.

By Lemma 1, $w_1 \leqslant q_1(t') \leqslant q_1(t'')$, and $q_2(t'') \leqslant q_2(t') \leqslant w_2$ almost everywhere for $0 \leqslant t' \leqslant t''$. Since $q_1$, $q_2$ are pointwise monotone, the limit functions in (4.3) and (4.4) exist on the manifold $X$. By construction, $q_1$, $q_{1,\infty}$, $q_2$ and $q_{2,\infty}$ obey (4.5).

In view of (4.5) and by the Lebesgue dominated convergence theorem, equalities (4.3) and (4.4) also hold in the norm $\|\,{\cdot}\,\|_{L^2(X)}$. The coefficients and right-hand side of equation (1.10) do not depend explicitly on $t$, and hence, by definitions (1.12) and (1.13), we have

$$ \begin{equation} \begin{aligned} \, &\langle q_j,p\rangle(t_0+t)-\langle q_j,p\rangle(t_0)+ \int_0^t\bigl(L(q_j,p)(t_0+\tau)- \langle q_j,p'\rangle(t_0+\tau)\bigr)\,d\tau \nonumber \\ &\qquad=\int_0^t\bigl(\langle f(\,\cdot\, ,q_j),p\rangle(t_0+\tau)+ \langle h,dp(t_0+\tau)\rangle_{L^2(T^* X)}\bigr)\,d\tau \end{aligned} \end{equation} \tag{4.6} $$
for $t_0 \geqslant 0$ and $p\in C^\infty([0,+\infty)\times X)$, and hence, for $0 \leqslant t' \leqslant t''$,
$$ \begin{equation*} \begin{aligned} \, &\langle q_j(t''+t)-q_j(t'+t),p(t)\rangle \\ &\quad\qquad+ \int_0^t\bigl(\mathcal{L}(q_j(t''+\tau)-q_j(t'+\tau),p(\tau)) -\langle q_j(t''+\tau)-q_j(t'+\tau),p'(\tau)\rangle\bigr)\,d\tau \\ &\quad=\langle q_j(t'')-q_j(t'),p(0)\rangle+ \int_0^t\langle f(\,\cdot\, ,q_j)(t''+\tau) -f(\,\cdot\, ,q_j)(t'+\tau),p(\tau)\rangle\,d\tau. \end{aligned} \end{equation*} \notag $$
Therefore, by Lemma 2, there exists a constant $C>0$ such that, for $t \in [0,+\infty)$,
$$ \begin{equation*} \begin{aligned} \, &\|q_j(t''+\,\cdot\,)-q_j(t'+\,\cdot\,)\|_{W(t;X)} \\ &\ \ \leqslant Ce^{Ct}\bigl(\|f(\,\cdot\, ,q_j)(t''+\,\cdot\,) -f(\,\cdot\, ,q_j)(t'+\,\cdot\,)\|_{L^2([0,t)\times X)}+ \|q_j(t'')-q_j(t')\|_{L^2(X)}\bigr). \end{aligned} \end{equation*} \notag $$
Hence, by the already established $\|\,{\cdot}\,\|_{L^2(X)}$-convergence (see (4.3) and (4.4)) and also by (4.6) and (1.5), we conclude that, as $t_0 \to +\infty$, the functions $q_1(t_0+\,\cdot\,)$ and $q_2(t_0+\,\cdot\,)$ satisfy the Cauchy condition in the norm $\|\,{\cdot}\,\|_{W(t;X)}$. By completeness of $W(t;X)$, they converge in this space to the functions $q_{1,\infty}$ (see (4.3)) and $q_{2,\infty}$ (see (4.4)), respectively. Choosing $p \in W^{1,2}(X)$ in (4.5) and making $t_0\to +\infty$, this gives
$$ \begin{equation*} \int_0^t \mathcal{L}(q_{j,\infty},p)\,d\tau= \int_0^t\bigl(\langle f(\,\cdot\, ,q_{j,\infty}),p\rangle+ \langle h,dp\rangle_{L^2(T^* X)}\bigr)\,d\tau \end{equation*} \notag $$
for $j=0,1$ and $t\in [0,T)$. The integrands in this equality are independent of $\tau$, and hence $q_{1,\infty}$ and $q_{2,\infty}$ are solutions of equation (1.6). By well-known regularity properties (see § 8.9 in [22] and § 2.2 in [10]), these solutions are continuous, and so the convergence in (4.4) also holds in the norm $\|\,{\cdot}\,\|_{C(X)}$ by Dini’s test. This proves Lemma 3.

4.2. Proof of Theorem 4

(a) By conditions of the theorem and in view of Remark 5, the numbers $r_0$ and $N$ are, respectively, a subsolution and a supersolution of problem (1.10), (1.11). Therefore, by Theorem 3, on the infinite interval $[0,+\infty)$ there exists a unique solution $q$ of problem (1.10), (1.11); moreover, $q \in C(+\infty;L^2(X))$, is locally Hölder-continuous on $(0,+\infty)\times X$, and $r_0 \leqslant q(t) \leqslant N$ for $t\in[0,+\infty)$. By definitions (1.12) and (2.1),

$$ \begin{equation*} \begin{aligned} \, &\langle r_0,p(t)\rangle+\mathcal{L}^t(r_0,p)-\langle r_0,p(0)\rangle- \int_0^t \langle f(\,\cdot\, ,r_0),p(\tau)\rangle\,d\tau \\ &\qquad \leqslant\langle q(t),p(t)\rangle+\mathcal{L}^t(q,p)- \langle q(0),p(0)\rangle- \int_0^t\langle f(\,\cdot\, ,q(\tau)),p(\tau)\rangle\,d\tau \end{aligned} \end{equation*} \notag $$
for all non-negative $p \in C^\infty([0,+\infty)\times X)$, which implies $q(t)>r_0$ for $t \in (0,+\infty)$ by Lemma 1. In particular, $q(1)>r_0$, and hence, by the conditions of the theorem, there exists $\delta_0>0$ such that $r_0+\alpha w$ is a subsolution of equation (1.6) satisfying $r_0+\alpha w \leqslant q(1)$ almost everywhere for $0<\alpha<\delta_0$.

By Lemma 3, for $w_1=r_0+\alpha w$ and $w_2=N$, there exist unique solutions $q_1$ and $q_2$ of the Cauchy problems (4.1) and (4.2) on the infinite interval $[0,+\infty)$; moreover, $q_1,q_2\in C(+\infty;L^2(X))$ and $q_1,q_2$ are locally Hölder-continuous on $(0,+\infty)\times X$. The $\|\,{\cdot}\,\|_{C(X)}$-limits $q_{1,\infty}$ (see (4.3)) and $q_{2,\infty}$ (see (4.4)) exist, satisfy equation (1.6), and obey (4.5) almost everywhere. From (4.5) it follows that $q_{1,\infty},q_{2,\infty}>r_0$ by the strong maximum principle for elliptic equations (see § 8.7 in [22]), and, therefore, by the uniqueness theorem, $q_{1,\infty}=q_{2,\infty}$ (see § 2.3 in [10]). Moreover, for $t \in [0,+\infty)$ and $0<\alpha<\delta_0$, by Lemma 1, we have $q_1(t) \leqslant q(t+1)$ and $q(t) \leqslant q_2(t)$, and hence, for $u=q_{1,\infty}=q_{1,\infty}$ and $\widetilde{q}(t)=\max\{u-q_1(t),q_2(t)-u\}$, we have (2.6), and, as a corollary, the limit relation (2.5).

Assertion (b) is verified similarly. Theorem 4 is proved.

4.3. Proof of Theorem 5

(a) By conditions of the theorem and in view of Remark 5, $r_0$ is a subsolution, and $r_0+Aw$ is a supersolution of problem (1.10), (1.11). Hence an appeal to Theorem 3 and Lemma 3 with $w_1=r_0$ and $w_2=r_0+Aw$ shows that there exist unique solutions $q$ and $q_2$ of problems (1.10), (1.11) and (4.2) on the infinite interval $[0,+\infty)$, where $q,q_2\in C(+\infty;L^2(X))$ and $q,q_2$ are locally Hölder-continuous on $(0,+\infty)\times X$. Next, by Lemma 3, the $\|\,{\cdot}\,\|_{C(X)}$-limit $q_{2,\infty}$ exists (see (4.4)) and is a solution of equation (1.6). Furthermore, by (4.5) $r_0 \leqslant q_{2,\infty} \leqslant r_0+Aw$ almost everywhere. Hence $q_{2,\infty}=r_0$ by the well-known uniqueness theorem (see § 2.3 in [10]). Next, by Lemma 1, $r_0 \leqslant q(t) \leqslant q_2(t)$ almost everywhere for $t\in [0,+\infty)$, and hence $\widetilde{q}(t)=q_2(t)-r_0$ obeys (2.6), and, therefore, and the limit relation (2.7).

Assertion (b) is proved similarly. Theorem 5 is proved.

§ 5. The eigenvalue problem

5.1. Proof of Theorem 7

(a) By Theorem 6, an eigenfunction $w \in C(X)$ belonging to the eigenvalue $\lambda_1$ can be chosen so as to have $\inf_{x \in X}w(x)\,{>}\,0$. By the assumption $\lambda_1<0$, and hence, by definition of the right derivative there exists $\delta>0$ such that, for all $0<\alpha<\delta$,

$$ \begin{equation*} \frac{\lambda_1}{2}\alpha w<f(\,\cdot\, ,r_0+\alpha w)- f(\,\cdot\, ,r_0)-\alpha w\,\frac{\partial f}{\partial r}(\,\cdot\, ,r_0+0) \end{equation*} \notag $$
almost everywhere. Now from (2.10), (2.9) we have
$$ \begin{equation*} \begin{aligned} \, 0&=\alpha\mathcal{L}(w,v)-\alpha\biggl\langle\frac{\partial f}{\partial r} (\,\cdot\, ,r_0+0)w,v\biggr\rangle-\lambda_1\alpha\langle w,v\rangle \\ &\geqslant \alpha\mathcal{L}(w,v)-\langle f(\,\cdot\, ,r_0+\alpha w),v\rangle+ \langle f(\,\cdot\, ,r_0),v\rangle+ \frac{\lambda_1}{2}\alpha\langle w,v\rangle-\lambda_1\alpha\langle w,v\rangle \end{aligned} \end{equation*} \notag $$
for non-negative functions $v \in C^\infty(X)$ (cf. § 3.2 in [3]). Therefore,
$$ \begin{equation*} \begin{aligned} \, &\alpha \mathcal{L}(w,v)+r_0\langle c,dv\rangle_{L^2(T^* X)}- \langle f(\,\cdot\, ,r_0+\alpha w),v\rangle- \langle h,dv\rangle_{L^2(T^* X)} \\ &\qquad\leqslant r_0\langle c,dv\rangle_{L^2(T^* X)}- \langle f(\,\cdot\, ,r_0),v\rangle-\langle h,dv\rangle_{L^2(T^* X)}+ \frac{\lambda_1}{2}\alpha\langle w,v\rangle. \end{aligned} \end{equation*} \notag $$
Hence by the definitions of $\mathcal{L}$ (see (1.8)) and $\theta$ (see (1.9)), for all $0<\alpha<\delta$, we have
$$ \begin{equation*} \mathcal{L}(r_0+\alpha w,v)-\langle f(\,\cdot\, ,r_0+\alpha w),v\rangle- \langle h,dv\rangle_{L^2(T^* X)} \leqslant \theta(r_0,v)+\frac{\lambda_1}{2}\alpha\langle w,v\rangle. \end{equation*} \notag $$
By Remark 5, a number $r_0$ is a subsolution of (1.6) if and only if $\theta(r_0,v) \leqslant 0$ for each non-negative function $v \in C^\infty(X)$. Next, since $\inf_{x \in X}w(x)>0$ and $\lambda_1<0$, and hence, for $0<\alpha<\delta$, the function $r_0+\alpha w$ is a subsolution of equation (1.6). So, the conditions of assertion (a) of Theorem 4 are met. Applying this assertion we get the required result. Assertion (b) is verified similarly. Theorem 7 is proved.

5.2. Proof of Theorem 8

(a) By Theorem 6, an eigenfunction $w \in C(X)$ belonging to the eigenvalue $\lambda_1$ can be chosen so as to have $\inf_{x\in X}w(x) > 0$. Hence, by monotonicity of the difference relations (2.11), for any $\alpha>0$, we have

$$ \begin{equation} f(\,\cdot\, ,r_0+\alpha w)-f(\,\cdot\, ,r_0)< \alpha w\,\frac{\partial f}{\partial r}(\,\cdot\, ,r_0+0) \end{equation} \tag{5.1} $$
almost everywhere. Now an appeal to (2.10), (2.9) shows that
$$ \begin{equation*} \begin{aligned} \, 0&=\alpha\mathcal{L}(w,v)- \alpha\biggl\langle\frac{\partial f}{\partial r} (\,\cdot\, ,r_0+0)w,v\biggr\rangle-\lambda_1\alpha\langle w,v\rangle \\ &\leqslant \alpha\mathcal{L}(w,v)-\langle f(\,\cdot\, ,r_0+\alpha w),v\rangle+ \langle f(\,\cdot\, ,r_0),v\rangle-\lambda_1\alpha\langle w,v\rangle \end{aligned} \end{equation*} \notag $$
for non-negative functions $v \in C^\infty(X)$ (cf. § 3.2 in [3]). Therefore,
$$ \begin{equation*} \begin{aligned} \, &\alpha\mathcal{L}(w,v)+r_0\langle c,dv\rangle_{L^2(T^* X)}- \langle f(\,\cdot\, ,r_0+\alpha w),v\rangle-\langle h,dv\rangle_{L^2(T^* X)} \\ &\qquad\geqslant r_0\langle c,dv\rangle_{L^2(T^* X)}- \langle f(\,\cdot\, ,r_0),v\rangle-\langle h,dv\rangle_{L^2(T^* X)}+ \lambda_1 \alpha\langle w,v\rangle, \end{aligned} \end{equation*} \notag $$
and now by definitions of $\mathcal{L}$ (see (1.8)) and $\theta$ (see (1.9)), for any $\alpha>0$, we have
$$ \begin{equation*} \mathcal{L}(r_0+\alpha w,v)-\langle f(\,\cdot\, ,r_0+\alpha w),v\rangle- \langle h,dv\rangle_{L^2(T^* X)} \geqslant \theta(r_0,v)+\lambda_1 \alpha\langle w,v\rangle. \end{equation*} \notag $$
In addition, since inequality (5.1) is strict, for $v=1$ we have
$$ \begin{equation*} \mathcal{L}(r_0+\alpha w,1)-\langle f(\,\cdot\, ,r_0+\alpha w),1\rangle> \theta(r_0,1)+\lambda_1\alpha\langle w,1\rangle. \end{equation*} \notag $$
By Remark 5, a number $r_0$ is a solution of equation (1.10) if and only if $\theta(r_0,v)=0$ for each non-negative function $v \in C^\infty(X)$. Since $\inf_{x\in X}w(x)>0$ and $\lambda_1\geqslant 0$, it follows that, for $\alpha>0$, the function $r_0+\alpha w$ is a strict supersolution of equation (1.6). Let $A>0$ be such that $r_0+Aw \geqslant q_0$ almost everywhere. This $A$ satisfies the conditions of assertion (a) of Theorem 5, which, in turn, secures the required result.

Assertion (b) is verified similarly. Theorem 8 is proved.

The author is grateful to A. A. Davydov for posing the problem and useful discussions.


Bibliography

1. A. N. Kolmogorov, I. G. Petrovskii, and N. S. Piskunov, The investigation of a diffusion equation connected with an increasing amount of substance and its application to a biological problem, Byull. Moskov. Univ., Mat. Mekh., no. 1, 1937 (Russian); French transl. A. Kolmogoroff, I. Petrovsky, N. Piscounoff, “Étude de l'équation de la diffusion avec croissance de la quantite de matière et son application à un problème biologique”, Moscou Univ. Bull. Math., 1:6 (1937), 1–25  zmath
2. R. A. Fisher, “The wave of advance of advantageous genes”, Ann. Eugenics, 7:4 (1937), 335–369  crossref  zmath
3. H. Berestycki, F. Hamel, and L. Roques, “Analysis of the periodically fragmented environment model. I. Species persistence”, J. Math. Biol., 51:1 (2005), 75–113  crossref  mathscinet  zmath
4. B. Perthame, Parabolic equations in biology. Growth, reaction, movement and diffusion, Lect. Notes Math. Model. Life Sci., Springer, Cham, 2015  crossref  mathscinet  zmath
5. G. Perelman, The entropy formula for the Ricci flow and its geometric applications, 2002, arXiv: math/0211159
6. G. Perelman, Ricci flow with surgery on three-manifoldsя, 2003, arXiv: math/0303109
7. G. Perelman, Finite extinction time for the solutions to the Ricci flow on certain three-manifoldsя, 2003, arXiv: math/0307245
8. D. M. DeTurck, “Deforming metrics in the direction of their Ricci tensors”, J. Diff. Geom., 18:1 (1983), 157–162  crossref  mathscinet  zmath
9. L. I. Nicolaescu, Lectures on the geometry of manifolds, 3rd ed., World Sci. Publ., Hackensack, NJ, 2021  crossref  mathscinet  zmath
10. D. V. Tunitsky, “On solvability of semilinear second-order elliptic equations on closed manifolds”, Izv. RAN. Ser. Mat., 86:5 (2022), 97–115  mathnet  crossref  mathscinet; English transl. Izv. Math., 86:5 (2022), 925–942  crossref
11. R. E. Showalter, Monotone operators in Banach space and nonlinear partial differential equations, Math. Surveys Monogr., 49, Amer. Math. Soc., Providence, RI, 1997  crossref  mathscinet  zmath
12. J. L. Lions, Équations différentielles opérationnelles et problèmes aux limites, Grundlehren Math. Wiss., 111, Springer-Verlag, Berlin–Göttingen–Heidelberg, 1961  mathscinet  zmath
13. L. Hörmander, The analysis of linear partial differential operators, v. II, Grundlehren Math. Wiss., 257, Differential operators with constant coefficients, Springer-Verlag, Berlin, 1983  crossref  mathscinet  zmath; Russian transl. Mir, Moscow, 1986  mathscinet  zmath
14. R. S. Palais, Seminar on the Atiyah–Singer index theorem, With contributions by M. F. Atiyah, A. Borel, E. E. Floyd, R. T. Seeley, W. Shih, R. Solovay, Ann. of Math. Stud., 57, Princeton Univ. Press, Princeton, NJ, 1965  mathscinet  zmath; Russian transl. Mir, Moscow, 1970  mathscinet  zmath
15. R. O. Wells, Jr., Differential analysis on complex manifolds, Prentice-Hall Series in Modern Analysis, Prentice-Hall, Inc., Englewood Cliffs, NJ, 1973  mathscinet  zmath; Russian transl. Mir, Moscow, 1976  mathscinet
16. G. M. Lieberman, Second order parabolic differential equations, World Sci. Publ., River Edge, NJ, 1996  crossref  mathscinet  zmath
17. Mingxin Wang, Nonlinear second order parabolic equations, CRC Press, Boca Ration, FL, 2021  crossref  zmath
18. L. C. Evans, Partial differential equations, Grad. Stud. Math., 19, 2nd ed., Amer. Math. Soc., Providence, RI, 2010  mathscinet  zmath
19. M. G. Kreĭn and M. A. Rutman, “Linear operators leaving invariant a cone in a Banach space”, Uspekhi Mat. Nauk, 3:1(23) (1948), 3–95 (Russian)  mathnet  mathscinet  zmath; English transl. Amer. Math. Soc. Translation, 26, Amer. Math. Soc., New York, 1950  mathscinet
20. J. Smoller, Shock waves and reaction-diffusion equations, Grundlehren Math. Wiss., 258, 2nd ed., Springer-Verlag, New York, 1994  crossref  mathscinet  zmath
21. P. Hess, Periodic-parabolic boundary value problems and positivity, Pitman Res. Notes Math. Ser., 247, Longman Scientific & Technical, Harlow; John Wiley & Sons, New York, 1991  mathscinet  zmath
22. D. Gilbarg and N. S. Trudinger, Elliptic partial differential equations of second order, Grundlehren Math. Wiss., 224, 2nd ed., Springer-Verlag, Berlin, 1983  crossref  mathscinet  zmath; Russian transl. Nauka, Moscow, 1989  mathscinet  zmath
23. R. Courant and D. Hilbert, Methods of mathematical physics, v. II, Partial differential equations, (vol. II by R. Courant), Interscience Publishers, a division of John Wiley & Sons, New York–London, 1962  mathscinet  zmath; Russian transl. Mir, Moscow, 1964  mathscinet  zmath
24. D. H. Sattinger, “Monotone methods in nonlinear elliptic and parabolic boundary value problems”, Indiana Univ. Math. J., 21:11 (1972), 979–1000  crossref  mathscinet  zmath

Citation: D. V. Tunitsky, “On stabilization of solutions of second-order semilinear parabolic equations on closed manifolds”, Izv. RAN. Ser. Mat., 87:4 (2023), 186–204; Izv. Math., 87:4 (2023), 817–834
Citation in format AMSBIB
\Bibitem{Tun23}
\by D.~V.~Tunitsky
\paper On stabilization of solutions of second-order semilinear parabolic equations on closed manifolds
\jour Izv. RAN. Ser. Mat.
\yr 2023
\vol 87
\issue 4
\pages 186--204
\mathnet{http://mi.mathnet.ru/im9354}
\crossref{https://doi.org/10.4213/im9354}
\mathscinet{http://mathscinet.ams.org/mathscinet-getitem?mr=4656043}
\adsnasa{https://adsabs.harvard.edu/cgi-bin/bib_query?2023IzMat..87..817T}
\transl
\jour Izv. Math.
\yr 2023
\vol 87
\issue 4
\pages 817--834
\crossref{https://doi.org/10.4213/im9354e}
\isi{https://gateway.webofknowledge.com/gateway/Gateway.cgi?GWVersion=2&SrcApp=Publons&SrcAuth=Publons_CEL&DestLinkType=FullRecord&DestApp=WOS_CPL&KeyUT=001088986700006}
\scopus{https://www.scopus.com/record/display.url?origin=inward&eid=2-s2.0-85174924187}
Linking options:
  • https://www.mathnet.ru/eng/im9354
  • https://doi.org/10.4213/im9354e
  • https://www.mathnet.ru/eng/im/v87/i4/p186
  • This publication is cited in the following 3 articles:
    Citing articles in Google Scholar: Russian citations, English citations
    Related articles in Google Scholar: Russian articles, English articles
    Известия Российской академии наук. Серия математическая Izvestiya: Mathematics
    Statistics & downloads:
    Abstract page:329
    Russian version PDF:10
    English version PDF:43
    Russian version HTML:75
    English version HTML:137
    References:63
    First page:9
     
      Contact us:
     Terms of Use  Registration to the website  Logotypes © Steklov Mathematical Institute RAS, 2024