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

HEAT EQUATION ON THE HYPERGRAPH CONTAINING VERTICES WITH GIVEN DATA

Takeshi Fukao, Masahiro Ikeda, Shun UchidaCorresponding author: shunuchida@oita-u.ac.jp
Abstract

This paper is concerned with the Cauchy problem of a multivalued ordinary differential equation governed by the hypergraph Laplacian, which describes the diffusion of “heat” or “particles” on the vertices of hypergraph. We consider the case where the heat on several vertices are manipulated internally by the observer, namely, are fixed by some given functions. This situation can be reduced to a nonlinear evolution equation associated with a time-dependent subdifferential operator, whose solvability has been investigated in numerous previous researches. In this paper, however, we give an alternative proof of the solvability in order to avoid some complicated calculations arising from the chain rule for the time-dependent subdifferential. As for results which cannot be assured by the known abstract theory, we also discuss the continuous dependence of solution on the given data and the time-global behavior of solution.

2020 Mathematics Subject Classification: Primary 34G25; Secondary 05C65, 47J30, 47J35. Keywords: Hypergraph, hypergraph Laplacian, subdifferential, nonlinear evolution equation, constraint problem.

1 Introduction

Let V={v1,,vn,vn+1,,vn+m}V=\{v_{1},\ldots,v_{n},v_{n+1},\ldots,v_{n+m}\} be a finite set, E2VE\subset 2^{V} be a family of subsets consisting of two or more elements of VV, and w:E(0,)w:E\to(0,\infty). The triplet G=(V,E,w)G=(V,E,w), which is called hypergraph, can be interpreted as a model of a network structure in which vertices v1,,vn+mVv_{1},\ldots,v_{n+m}\in V are connected by each hyperedge eEe\in E (see Fig. 1).

Refer to caption
Refer to caption
Figure 1: When #e=2e=2 for every eEe\in E, each edge eEe\in E can be regarded as a segment connecting two vertices (left figure, called usual graph). Hypergraph is a generalization of usual graph which represents the connection and grouping of multiple members (right figure).

In order to investigate the structure of network represented by a hypergraph, Yoshida [16] introduced a nonlinear set-valued operator on V\mathbb{R}^{V} (the family of mappings x:Vx:V\to\mathbb{R}) as follows. Define fe(x):=maxu,ve(x(u)x(v))f_{e}(x):=\max_{u,v\in e}(x(u)-x(v)) with respect to each eEe\in E and

(1) φG,p(x):=1peEw(e)(fe(x))pp[1,).\varphi_{G,p}(x):=\frac{1}{p}\sum_{e\in E}w(e)(f_{e}(x))^{p}~{}~{}~{}~{}~{}p\in[1,\infty).

These are convex functionals with domains D(fe)=D(φG,p)=VD(f_{e})=D(\varphi_{G,p})=\mathbb{R}^{V} and then subdifferentiable at every xVx\in\mathbb{R}^{V}. By the chain rule (see, e.g., [2]) and the maximum rule (see, e.g., [13, Proposition 2.54]), the subdifferential of φG,p\varphi_{G,p} coincides with

(2) φG,p(x)={eEw(e)(fe(x))p1be;beargmaxbBebx},\partial\varphi_{G,p}(x)=\left\{\sum_{e\in E}w(e)(f_{e}(x))^{p-1}b_{e};~{}b_{e}\in\operatorname*{argmax}_{b\in B_{e}}b\cdot x\right\},

where BeVB_{e}\subset\mathbb{R}^{V}, called base polytope for eEe\in E, is defined by

(3) Be:=conv{1u1v;u,ve}B_{e}:=\operatorname*{conv}\{1_{u}-1_{v};~{}u,v\in e\}

and 1u:V1_{u}:V\to\mathbb{R} with uVu\in V stands for

(4) 1u(v):={1 if u=v,0 if uv.1_{u}(v):=\begin{cases}~{}~{}1~{}~{}&~{}~{}\text{ if }u=v,\\ ~{}~{}0~{}~{}&~{}~{}\text{ if }u\neq v.\end{cases}

This operator LG,p=φG,p:V2VL_{G,p}=\partial\varphi_{G,p}:\mathbb{R}^{V}\to 2^{\mathbb{R}^{V}} is called hypergraph (pp-)Laplacian, where 1p<1\leq p<\infty (see also [7]).

When p=2p=2 and every eEe\in E consists of two elements of VV, then LG,2L_{G,2} becomes a linear operator on V\mathbb{R}^{V}, i.e., a square matrix of order #V=(n+m){\#}V=(n+m) and represents the movement of “particles” on a vertex to another adjacent vertex through an edge eEe\in E in the random walk model on the graph. By analogy with this, we can regard the ODE x(t)+LG,p(x(t))0x^{\prime}(t)+L_{G,p}(x(t))\ni 0 with respect to x:[0,T]Vx:[0,T]\to\mathbb{R}^{V} as a diffusion model of “heat” or “particles” on each vertex viv_{i}, which is described by x(t)(vi)=xi(t)x(t)(v_{i})=x_{i}(t). In fact, the time-global behavior of solution to this ODE (studied in [7]) quite resembles those to the PDE tu(|u|p2u)=0\partial_{t}u-\nabla\cdot(|\nabla u|^{p-2}\nabla u)=0. In information science, this ODE x(t)+LG,p(x(t))0x^{\prime}(t)+L_{G,p}(x(t))\ni 0 is used to investigate the eigenvalues of LG,pL_{G,p}, prove the Cheeger like inequality, and analyze the cluster structure of network represented by hypergraph (see, e.g., [3, 6, 12, 15]).

Here let the heat on two vertices vi,vjv_{i},v_{j} be fixed as x(vi)=αx(v_{i})=\alpha and x(vj)=βx(v_{j})=\beta (α>β\alpha>\beta). Under this situation, the final state x=limtx(t)x_{\infty}=\lim_{t\to\infty}x(t) of the diffusion process x(t)+LG,p(x(t))0x^{\prime}(t)+L_{G,p}(x(t))\ni 0 might describe the thermal gradient in each path and then indicate the optimal path of heat flux from vertex viv_{i} to vjv_{j}. As a generalization of this problem, we deal with the case where the heat at vertices vn+1,,vn+mv_{n+1},\ldots,v_{n+m} are determined by some given functions aj:[0,T]a_{j}:[0,T]\to\mathbb{R} (j=1,,mj=1,\ldots,m), i.e., the heat equation governed by LG,pL_{G,p} under the condition

(5) xn+j(t)=x(t)(vn+j)=aj(t)j=1,,mx_{n+j}(t)=x(t)(v_{n+j})=a_{j}(t)~{}~{}~{}~{}~{}j=1,\ldots,m

(see Fig. 2). This can be interpreted as a situation where the observer internally manipulates the heat of the network. As another example, the graph Laplacian (2) and its energy functional (1) on the usual graph can be found in the definition of Laplacian on the fractal (see, e.g., [10, 11]). In this case, the assumption aj0a_{j}\equiv 0 corresponds to a homogeneous Dirichlet boundary condition on the fractal.

Refer to caption
Figure 2: We consider the case where the “heat” of some vertices (colored in the above) is given as xn+i(t)=ai(t)x_{n+i}(t)=a_{i}(t).

We here state the setting of our problem more precisely. Without loss of generality, we assume the hypergraph G=(V,E,w)G=(V,E,w) is connected, i.e., for every u,vVu,v\in V there exist some u1,uN1Vu_{1},\ldots u_{N-1}\in V and e1,e2,eNEe_{1},e_{2},\ldots e_{N}\in E such that uj1,ujeju_{j-1},u_{j}\in e_{j} holds for any j=1,2,,Nj=1,2,\ldots,N, where u0=uu_{0}=u and uN=vu_{N}=v (if GG is disconnected, we only have to divide GG into connected components and consider the problem on each component, see [7]).

By letting xi:=x(vi)x_{i}:=x(v_{i}), we can identify V\mathbb{R}^{V} (the family of mappings from VV onto \mathbb{R}) with the Euclidean space n+m\mathbb{R}^{n+m} ((4) coincides with a unit vector of the canonical basis). In this sense, V\mathbb{R}^{V} is a Hilbert space endowed with the standard inner product xy:=vVx(v)y(v)x\cdot y:=\sum_{v\in V}x(v)y(v) and the norm x:=vVx(v)2\|x\|:=\sqrt{\sum_{v\in V}x(v)^{2}}. For x:[0,T]Vx:[0,T]\to\mathbb{R}^{V}, we shall write xi(t):=x(t)(vi)x_{i}(t):=x(t)(v_{i}) for the same reason.

For convenience, the given data in (5) will be unified by a:[0,T]Va:[0,T]\to\mathbb{R}^{V} as

(6) a(t):=(0,,0,a1(t),,am(t)).a(t):=(0,\ldots,0,a_{1}(t),\ldots,a_{m}(t)).

Let Ka(t)VK_{a}(t)\subset\mathbb{R}^{V} be a constraint set which corresponds to the condition (5) such that the former nn components x1,,xnx_{1},\ldots,x_{n} are chosen freely and the latter mm components xn+1,,xn+mx_{n+1},\ldots,x_{n+m} are fixed by given data a1(t),,am(t)a_{1}(t),\ldots,a_{m}(t). Namely, define

Ka(t):={xV;x=(x1,,xn,a1(t),,am(t)),xi(i=1,,n)}.K_{a}(t):=\{x\in\mathbb{R}^{V};~{}x=(x_{1},\ldots,x_{n},a_{1}(t),\ldots,a_{m}(t)),~{}x_{i}\in\mathbb{R}~{}(i=1,\ldots,n)\}.

Note that Ka(t)K_{a}(t) is a closed convex subset in V\mathbb{R}^{V} and the projection onto Ka(t)K_{a}(t) can be defined by

(7) ProjKa(t)x:=(x1,,xn,a1(t),,am(t)),where x=(x1,,xn+m).\text{Proj}_{K_{a}(t)}x:=(x_{1},\ldots,x_{n},a_{1}(t),\ldots,a_{m}(t)),~{}~{}~{}\text{where ~{}}x=(x_{1},\ldots,x_{n+m}).

Moreover, let IKa(t):V[0,]I_{K_{a}(t)}:\mathbb{R}^{V}\to[0,\infty] stand for the indicator on Ka(t)VK_{a}(t)\subset\mathbb{R}^{V}:

(8) IKa(t)(x)={0 if xKa(t),+ otherwise.I_{K_{a}(t)}(x)=\begin{cases}~{}~{}0~{}~{}&~{}~{}\text{ if }x\in K_{a}(t),\\ ~{}~{}+\infty~{}~{}&~{}~{}\text{ otherwise.}\end{cases}

We can define the subdifferential of IKa(t)I_{K_{a}(t)} on Ka(t)K_{a}(t). Here when ψ:V(,+]\psi:\mathbb{R}^{V}\to(-\infty,+\infty] is a proper (ψ+\psi\not\equiv+\infty) lower semi-continuous and convex function, its subgradient at xVx\in\mathbb{R}^{V} is defined by

(9) ψ(x):={ηV;η(zx)ψ(z)ψ(x)zV}.\partial\psi(x):=\{\eta\in\mathbb{R}^{V};~{}\eta\cdot(z-x)\leq\psi(z)-\psi(x)~{}~{}\forall z\in\mathbb{R}^{V}\}.

Then the diffusion process on the hypergraph with the condition (5) and the initial state x0Vx_{0}\in\mathbb{R}^{V} can be represented by the following Cauchy problem of an evolution equation with constraint condition:

(P)a,h,x0{x(t)+φG,p(x(t))+IKa(t)(x(t))h(t)t[0,T],x(0)=x0,\text{(P)}_{a,h,x_{0}}~{}~{}~{}~{}~{}~{}~{}\begin{cases}~{}~{}\displaystyle x^{\prime}(t)+\partial\varphi_{G,p}(x(t))+\partial I_{K_{a}(t)}(x(t))\ni h(t)~{}~{}~{}t\in[0,T],\\ ~{}~{}x(0)=x_{0},\end{cases}

where h:[0,T]Vh:[0,T]\to\mathbb{R}^{V} is the given external force.

Since φG,p(x)<\varphi_{G,p}(x)<\infty holds for every xVx\in\mathbb{R}^{V} (i.e., D(φG,p)=VD(\varphi_{G,p})=\mathbb{R}^{V}), we have (φG,p+IKa(t))=φG,p+IKa(t)\partial(\varphi_{G,p}+I_{K_{a}(t)})=\partial\varphi_{G,p}+\partial I_{K_{a}(t)}. Therefore (P)a,h,x0{}_{a,h,x_{0}} can be reduced to a problem of evolution equation governed by a subdifferential operator of time-dependent functional φt:=φG,p+IKa(t)\varphi^{t}:=\varphi_{G,p}+I_{K_{a}(t)}, which has been investigated in numerous papers, e.g., [4, 5, 8, 9, 14]. In this paper, however, we aim to introduce a simpler method for the constraint problem (P)a,h,x0{}_{a,h,x_{0}} by focusing on features of φG,p\partial\varphi_{G,p} and IKa(t)\partial I_{K_{a}(t)} stated in the next section. We prove the existence of solution to (P)a,h,x0{}_{a,h,x_{0}} in §3. As for a result which cannot be assured by the known abstract theory, we discuss the continuous dependence of solution on the given data a(t)a(t) in §4. Finally, we consider the time-global behavior of solution in §5 by using the facts given in previous parts.

2 Preliminary

In this section, we check and review some basic facts for later use. First, the subgradient of (8) can be written specifically as follows:

Lemma 2.1.

For any t[0,T]t\in[0,T] and xKa(t)x\in K_{a}(t), the subdifferential of IKa(t)I_{K_{a}(t)} is characterized by

IKa(t)(x)={(0,0,ξn+1,,ξn+m)V,ξn+j(j=1,,m)}.\partial I_{K_{a}(t)}(x)=\{(0\ldots,0,\xi_{n+1},\ldots,\xi_{n+m})\in\mathbb{R}^{V},~{}\xi_{n+j}\in\mathbb{R}~{}~{}(j=1,\ldots,m)\}.
Proof..

Let ξIKa(t)(x)\xi\in\partial I_{K_{a}(t)}(x). By the definition of subdifferntial (9), 0ξ(zx)0\geq\xi\cdot(z-x) holds for every zKa(t)z\in K_{a}(t). When x,zKa(t)x,z\in K_{a}(t), we have

zx\displaystyle z-x =(z1x1,,znxn,a1(t)a1(t),,am(t)am(t))\displaystyle=(z_{1}-x_{1},\ldots,z_{n}-x_{n},a_{1}(t)-a_{1}(t),\ldots,a_{m}(t)-a_{m}(t))
=(z1x1,,znxn,0,,0).\displaystyle=(z_{1}-x_{1},\ldots,z_{n}-x_{n},0,\ldots,0).

Hence it is necessary that 0i=1nξi(zixi)0\geq\sum_{i=1}^{n}\xi_{i}(z_{i}-x_{i}) holds for any ziz_{i}\in\mathbb{R} (i=1,,ni=1,\ldots,n). This implies that ξ1,,ξn\xi_{1},\ldots,\xi_{n} should be equal to zero and ξn+1,,ξn+m\xi_{n+1},\ldots,\xi_{n+m} can be chosen arbitrarily. ∎

We next check the following Poincaré type inequality by repeating the argument in [7]. Here and henceforth, φG,p\varphi_{G,p} in (1) will be abbreviated to φ\varphi.

Theorem 2.2.

Let the hypergraph G=(V,E,w)G=(V,E,w) be connected and xKa(t)x\in K_{a}(t). Then there exists some constant CC depending only on pp and GG such that

(10) k=1n|xk|Cφ(x)1/p+nmin1im|ai(t)|.\sum_{k=1}^{n}|x_{k}|\leq C\varphi(x)^{1/p}+n\min_{1\leq i\leq m}|a_{i}(t)|.
Proof..

Let the index jj satisfy |aj(t)|=min1im|ai(t)||a_{j}(t)|=\min_{1\leq i\leq m}|a_{i}(t)|. Since GG is assumed to be connected, for every fixed vkv_{k} (k=1,,nk=1,\ldots,n) there exist some u1,,uN1Vu_{1},\ldots,u_{N-1}\in V and e1,e2,,eNEe_{1},e_{2},\ldots,e_{N}\in E s.t. ul1,ulelu_{l-1},u_{l}\in e_{l} (l=1,2,,Nl=1,2,\ldots,N) holds, where u0=vku_{0}=v_{k} and uN=vn+ju_{N}=v_{n+j}. By x(vn+j)=aj(t)x(v_{n+j})=a_{j}(t) and Hölder’s inequality, we have

|x(vk)aj(t)|\displaystyle|x(v_{k})-a_{j}(t)| l=1N|x(ul1)x(ul)|l=1Nfel(x)eEfe(x)\displaystyle\leq\sum_{l=1}^{N}|x(u_{l-1})-x(u_{l})|\leq\sum_{l=1}^{N}f_{e_{l}}(x)\leq\sum_{e\in E}f_{e}(x)
(#V)1/p(mineEw(e))1/pφ(x)1/p,\displaystyle\leq\frac{({\#}V)^{1/p^{\prime}}}{(\min_{e\in E}w(e))^{1/p}}\varphi(x)^{1/p},

where p:=p/(p1)p^{\prime}:=p/(p-1) is the Hölder conjugate exponent. Then we obtain (10) with the positive constant C=n(n+m)1/p(mineEw(e))1/pC=n(n+m)^{1/p^{\prime}}(\min_{e\in E}w(e))^{-1/p}. ∎

Note that (see [7, Theorem 2.4])

(11) φ(y)=0cs.t.y=(c,,c).\varphi(y)=0~{}~{}\Leftrightarrow~{}~{}\exists c\in\mathbb{R}~{}~{}\text{s.t.}~{}~{}y=(c,\ldots,c).

In addition, we can easily show that fe(x)=maxu,ve|x(u)x(v)|2xf_{e}(x)=\max_{u,v\in e}|x(u)-x(v)|\leq 2\|x\| by definition and b2\|b\|\leq 2 for any bBeb\in B_{e} since 1uV1_{u}\in\mathbb{R}^{V} can be identified with a unit vector of n+m\mathbb{R}^{n+m}. Hence as for the boundedness of φ\varphi and φ\partial\varphi,

(12) φ(x)2p(#E)maxeEw(e)pxp,η2p(#E)maxeEw(e)xp1\varphi(x)\leq\frac{2^{p}({\#}E)\max_{e\in E}w(e)}{p}\|x\|^{p},\hskip 14.22636pt\|\eta\|\leq 2^{p}({\#}E)\max_{e\in E}w(e)\|x\|^{p-1}

hold for every xVx\in\mathbb{R}^{V} and ηφ(x)\eta\in\partial\varphi(x), where #E{\#}E is the number of hyperedges.

We also recall a variant of Gronwall’s inequality as follows (see [1, Lemme A.5]). Let Lq(0,T;V)L^{q}(0,T;\mathbb{R}^{V}) (q[1,)q\in[1,\infty)) and L(0,T;V)L^{\infty}(0,T;\mathbb{R}^{V}) denote the family of g:[0,T]Vg:[0,T]\to\mathbb{R}^{V} which satisfies 0Tg(t)q𝑑t<\int_{0}^{T}\|g(t)\|^{q}dt<\infty and sup0tTg(t)<\sup_{0\leq t\leq T}\|g(t)\|<\infty, respectively. Moreover, gW1,q(0,T;V)g\in W^{1,q}(0,T;\mathbb{R}^{V}) implies that gg and its derivative gg^{\prime} belong to Lq(0,T;V)L^{q}(0,T;\mathbb{R}^{V}).

Lemma 2.3.

Let Ξ:[0,T]\Xi:[0,T]\to\mathbb{R} be continuous and satisfy

12Ξ(t)212κ2+0tg(s)Ξ(s)𝑑st[0,T]\frac{1}{2}\Xi(t)^{2}\leq\frac{1}{2}\kappa^{2}+\int_{0}^{t}g(s)\Xi(s)ds~{}~{}~{}\forall t\in[0,T]

with some constant κ\kappa and non-negative function gL1(0,T;)g\in L^{1}(0,T;\mathbb{R}). Then Ξ\Xi satisfies

|Ξ(t)||κ|+0tg(s)𝑑st[0,T].|\Xi(t)|\leq|\kappa|+\int_{0}^{t}g(s)ds~{}~{}~{}\forall t\in[0,T].

3 Existence of Solution

We begin with the solvability of (P)a,h,x0{}_{a,h,x_{0}}:

Theorem 3.1.

Let T(0,)T\in(0,\infty) and aW1,2(0,T;V)a\in W^{1,2}(0,T;\mathbb{R}^{V}). Then for any x0Ka(0)x_{0}\in K_{a}(0) and hL2(0,T;V)h\in L^{2}(0,T;\mathbb{R}^{V}), (P)a,h,x0{}_{a,h,x_{0}} possesses a unique solution xW1,2(0,T;V)x\in W^{1,2}(0,T;\mathbb{R}^{V}).

This result can be assured by known results for abstract theory of nonlinear evolution equations with time-dependent subdifferential. Indeed, we can apply [9, Theorem 1.1.2] since φt:=φ+IKa(t)\varphi^{t}:=\varphi+I_{K_{a}(t)} satisfies the condition (H)2 in [9, §1.5] if aW1,2(0,T;V)a\in W^{1,2}(0,T;\mathbb{R}^{V}). In this section, however, we shall give an alternative proof which only relies on the basic properties of subdifferential in order to avoid some complicated calculations arising from the chain rule of ddtφt(x(t))\frac{d}{dt}\varphi^{t}(x(t)).

Proof..

Remark that the Moreau-Yosida regularization of IKa(t)I_{K_{a}(t)} and its subdifferential (the Yosida approximation of IKa(t)\partial I_{K_{a}(t)}) coincide with

(IKa(t))λ(x)\displaystyle(I_{K_{a}(t)})_{\lambda}(x) :=infyV{12λxy2+IKa(t)(y)}=12λxProjKa(t)x2,\displaystyle:=\inf_{y\in\mathbb{R}^{V}}\left\{\frac{1}{2\lambda}\|x-y\|^{2}+I_{K_{a}(t)}(y)\right\}=\frac{1}{2\lambda}\|x-\text{Proj}_{K_{a}(t)}x\|^{2},
(IKa(t))λ(x)\displaystyle(\partial I_{K_{a}(t)})_{\lambda}(x) :=(IKa(t))λ(x)=1λ(xProjKa(t)x)λ>0.\displaystyle:=\partial(I_{K_{a}(t)})_{\lambda}(x)=\frac{1}{\lambda}(x-\text{Proj}_{K_{a}(t)}x)\hskip 28.45274pt\lambda>0.

We first replace IKa(t)\partial I_{K_{a}(t)} in (P)a,h,x0{}_{a,h,x_{0}} with (IKa(t))λ(\partial I_{K_{a}(t)})_{\lambda} and consider the following approximation problem:

(P)λ{xλ(t)+φ(xλ(t))+xλ(t)ProjKa(t)xλ(t)λh(t),xλ(0)=x0.\text{(P)}_{\lambda}~{}~{}~{}~{}\begin{cases}~{}~{}\displaystyle x^{\prime}_{\lambda}(t)+\partial\varphi(x_{\lambda}(t))+\frac{x_{\lambda}(t)-\operatorname*{Proj}_{K_{a}(t)}x_{\lambda}(t)}{\lambda}\ni h(t),\\[8.53581pt] ~{}~{}x_{\lambda}(0)=x_{0}.\end{cases}

By the Lipschitz continuity of 1λ(xProjKa(t)x)\frac{1}{\lambda}(x-\operatorname*{Proj}_{K_{a}(t)}x), the abstract theory (see e.g, [1, Théorème 3.6]) yields the existence of a unique solution xλW1,2(0,T;V)x_{\lambda}\in W^{1,2}(0,T;\mathbb{R}^{V}) to (P)λ for every x0Vx_{0}\in\mathbb{R}^{V} and hL2(0,T;V)h\in L^{2}(0,T;\mathbb{R}^{V}).

We establish a priori estimates independent of λ\lambda. Since

xλ(t)ProjKa(t)xλ(t)=(0,,0,xλ,n+1(t)a1(t),,xλ,n+m(t)am(t)),x_{\lambda}(t)-\text{Proj}_{K_{a}(t)}x_{\lambda}(t)=(0,\ldots,0,x_{\lambda,n+1}(t)-a_{1}(t),\ldots,x_{\lambda,n+m}(t)-a_{m}(t)),

multiplying

(xλ(t)a(t))+φ(xλ(t))+xλ(t)ProjKa(t)xλ(t)λh(t)a(t)(x_{\lambda}(t)-a(t))^{\prime}+\partial\varphi(x_{\lambda}(t))+\frac{x_{\lambda}(t)-\operatorname*{Proj}_{K_{a}(t)}x_{\lambda}(t)}{\lambda}\ni h(t)-a^{\prime}(t)

by xλax_{\lambda}-a and using (9), we have

12ddtxλ(t)a(t)2+φ(xλ(t))φ(a(t))+1λj=1m|xλ,n+j(t)aj(t)|2\displaystyle\frac{1}{2}\frac{d}{dt}\|x_{\lambda}(t)-a(t)\|^{2}+\varphi(x_{\lambda}(t))-\varphi(a(t))+\frac{1}{\lambda}\sum_{j=1}^{m}|x_{\lambda,n+j}(t)-a_{j}(t)|^{2}
h(t)a(t)xλ(t)a(t).\displaystyle\hskip 85.35826pt\leq\|h(t)-a^{\prime}(t)\|\|x_{\lambda}(t)-a(t)\|.

Hence if xλ(0)=x0=(x01,,x0n,a1(0),,an(0))Ka(0)x_{\lambda}(0)=x_{0}=(x_{01},\ldots,x_{0n},a_{1}(0),\ldots,a_{n}(0))\in K_{a}(0), we obtain

(13) sup0tTxλ(t)a(t)2+2λ0Tj=1m|xλ,n+j(t)aj(t)|2dt(i=1n|x0i|2+20Tφ(a(t))𝑑t+0Th(t)a(t)2𝑑t)expT.\begin{split}&\sup_{0\leq t\leq T}\|x_{\lambda}(t)-a(t)\|^{2}+\frac{2}{\lambda}\int_{0}^{T}\sum_{j=1}^{m}|x_{\lambda,n+j}(t)-a_{j}(t)|^{2}dt\\ &\leq\left(\sum_{i=1}^{n}|x_{0i}|^{2}+2\int_{0}^{T}\varphi(a(t))dt+\int_{0}^{T}\|h(t)-a^{\prime}(t)\|^{2}dt\right)\exp{T}.\end{split}

From this and (12), there exists some constant C1C_{1} independent of λ\lambda such that

(14) sup0tTxλ(t)C1,sup0tTφ(xλ(t))+sup0tTηλ(t)C1,\begin{split}\sup_{0\leq t\leq T}\|x_{\lambda}(t)\|\leq C_{1},~{}~{}~{}\sup_{0\leq t\leq T}\varphi(x_{\lambda}(t))+\sup_{0\leq t\leq T}\|\eta_{\lambda}(t)\|\leq C_{1},\end{split}

where ηλ\eta_{\lambda} is the section of (P)λ, i.e., fulfills

xλ(t)+ηλ(t)+1λ(xλ(t)ProjKa(t)xλ(t))=h(t),ηλ(t)φ(xλ(t))x^{\prime}_{\lambda}(t)+\eta_{\lambda}(t)+\frac{1}{\lambda}(x_{\lambda}(t)-\text{Proj}_{K_{a}(t)}x_{\lambda}(t))=h(t),~{}~{}~{}~{}\eta_{\lambda}(t)\in\partial\varphi(x_{\lambda}(t))

for a.e. t[0,T]t\in[0,T]. We next test (P)λ by (xλ(t)a(t))(x_{\lambda}(t)-a(t))^{\prime}. Since φ\varphi does not depend on the time, the standard chain rule (ηλ(t),xλ(t))=ddtφ(xλ(t))(\eta_{\lambda}(t),x^{\prime}_{\lambda}(t))=\frac{d}{dt}\varphi(x_{\lambda}(t)) is valid (see [1, Lemme 3.3]). Moreover, from

(xλ(t)ProjKa(t)xλ(t),xλ(t)a(t))\displaystyle(x_{\lambda}(t)-\text{Proj}_{K_{a}(t)}x_{\lambda}(t),x^{\prime}_{\lambda}(t)-a^{\prime}(t))
=j=1m(xλ,n+j(t)aj(t))(xλ,n+j(t)aj(t))=12ddtj=1m|xλ,n+j(t)aj(t)|2\displaystyle=\sum_{j=1}^{m}(x_{\lambda,n+j}(t)-a_{j}(t))(x_{\lambda,n+j}(t)-a_{j}(t))^{\prime}=\frac{1}{2}\frac{d}{dt}\sum_{j=1}^{m}|x_{\lambda,n+j}(t)-a_{j}(t)|^{2}

and (14), we can derive

12xλ(t)a(t)2+ddtφ(xλ(t))+12λddtj=1m|xλ,n+j(t)aj(t)|2\displaystyle\frac{1}{2}\|x^{\prime}_{\lambda}(t)-a^{\prime}(t)\|^{2}+\frac{d}{dt}\varphi(x_{\lambda}(t))+\frac{1}{2\lambda}\frac{d}{dt}\sum_{j=1}^{m}|x_{\lambda,n+j}(t)-a_{j}(t)|^{2}
C1a(t)+12h(t)a(t)2.\displaystyle\leq C_{1}\|a^{\prime}(t)\|+\frac{1}{2}\|h(t)-a^{\prime}(t)\|^{2}.

By xλ,n+j(0)=aj(0)x_{\lambda,n+j}(0)=a_{j}(0), we get

(15) 0Txλ(t)a(t)2𝑑t+1λsup0tTj=1m|xλ,n+j(t)aj(t)|22φ(x0)+2C10Ta(t)𝑑t+0Th(t)a(t)2𝑑t.\begin{split}&\int_{0}^{T}\|x^{\prime}_{\lambda}(t)-a^{\prime}(t)\|^{2}dt+\frac{1}{\lambda}\sup_{0\leq t\leq T}\sum_{j=1}^{m}|x_{\lambda,n+j}(t)-a_{j}(t)|^{2}\\ &\leq 2\varphi(x_{0})+2C_{1}\int_{0}^{T}\|a^{\prime}(t)\|dt+\int_{0}^{T}\|h(t)-a^{\prime}(t)\|^{2}dt.\end{split}

By (14) and (15), we can apply the Ascoli-Arzela theorem and extract a subsequence of {xλ}λ>0\{x_{\lambda}\}_{\lambda>0} which converges uniformly in [0,T][0,T] (we omit relabeling since the original sequence also converges as will be seen later). Let xC([0,T];V)x\in C([0,T];\mathbb{R}^{V}) be its limit. Furthermore, (14) and (15) also lead to

ηλη,xλxweakly in L2(0,T;V)\eta_{\lambda}\rightharpoonup\exists\eta,~{}~{}~{}x^{\prime}_{\lambda}\rightharpoonup x^{\prime}~{}~{}~{}\text{weakly in }L^{2}(0,T;\mathbb{R}^{V})

and the demiclosedness of maximal monotone operators implies that the limit η\eta satisfies η(t)φ(x(t))\eta(t)\in\partial\varphi(x(t)) for a.e. t(0,T)t\in(0,T). Since (15) yields

sup0tTj=1m|xλ,n+j(t)aj(t)|2λC1,\sup_{0\leq t\leq T}\sum_{j=1}^{m}|x_{\lambda,n+j}(t)-a_{j}(t)|^{2}\leq\lambda C_{1},

we have xn+j(t)=aj(t)x_{n+j}(t)=a_{j}(t), i.e., x(t)Ka(t)x(t)\in K_{a}(t) for every t[0,T]t\in[0,T]. By the equation,

xλProjKa()xλλξweakly in L2(0,T;V)\frac{x_{\lambda}-\operatorname*{Proj}_{K_{a}(\cdot)}x_{\lambda}}{\lambda}\rightharpoonup\exists\xi~{}~{}~{}\text{weakly in }L^{2}(0,T;\mathbb{R}^{V})

and h(t)xλ(t)ηλ(t)=(IKa(t))λ(xλ(t))h(t)-x^{\prime}_{\lambda}(t)-\eta_{\lambda}(t)=\partial(I_{K_{a}(t)})_{\lambda}(x_{\lambda}(t)) (a.e. t[0,T]t\in[0,T]) hold. Then it follows from (9) that for every vL2(0,T;V)v\in L^{2}(0,T;\mathbb{R}^{V}) satisfying v(t)Ka(t)v(t)\in K_{a}(t),

0T(h(t)xλ(t)ηλ(t),v(t)xλ(t))𝑑t\displaystyle\int_{0}^{T}(h(t)-x^{\prime}_{\lambda}(t)-\eta_{\lambda}(t),v(t)-x_{\lambda}(t))dt
0T(IKa(t))λ(v(t))𝑑t0T(IKa(t))λ(xλ(t))𝑑t0.\displaystyle\leq\int_{0}^{T}(I_{K_{a}(t)})_{\lambda}(v(t))dt-\int_{0}^{T}(I_{K_{a}(t)})_{\lambda}(x_{\lambda}(t))dt\leq 0.

By taking its limit as λ0\lambda\to 0, we obtain

0T(h(t)x(t)η(t),v(t)x(t))𝑑t0vL2(0,T;V)withv(t)Ka(t),\int_{0}^{T}(h(t)-x^{\prime}(t)-\eta(t),v(t)-x(t))dt\leq 0~{}~{}~{}~{}\forall v\in L^{2}(0,T;\mathbb{R}^{V})~{}~{}\text{with}~{}~{}v(t)\in K_{a}(t),

which entails ξ(t)=h(t)x(t)η(t)IKa(t)(x(t))\xi(t)=h(t)-x^{\prime}(t)-\eta(t)\in\partial I_{K_{a}(t)}(x(t)) a.e. t[0,T]t\in[0,T]. Thus the limit xx fulfills all requirements of the solution to (P)a,h,x0{}_{a,h,x_{0}}.

The uniqueness of solution can be guaranteed by the monotonicity of φ\partial\varphi and IKa(t)\partial I_{K_{a}(t)}. Therefore the limit xx is determined independently of the choice of subsequences and then the original sequence {xλ}\{x_{\lambda}\} also converges to xx. ∎

Remark 3.2.

We can find several abstract results for the case where the movement of the constraint set is not smooth (see, e.g., [4, §4]). However, it seems to be difficult to mitigate our assumption aW1,2(0,T;V)a\in W^{1,2}(0,T;\mathbb{R}^{V}) since Ka(t)K_{a}(t) does not possess any interior point and the smoothness of movement of Ka(t)K_{a}(t) must exactly coincide with the regularity of the solution.

4 Continuous Dependence of Solutions on Given data

We next consider the dependence of solutions with respect to the given data.

Theorem 4.1.

Let (ak,hk,x0k)W1,2(0,T;V)×L2(0,T;V)×Kak(0)(a^{k},h^{k},x^{k}_{0})\in W^{1,2}(0,T;\mathbb{R}^{V})\times L^{2}(0,T;\mathbb{R}^{V})\times K_{a^{k}}(0) and let xkW1,2(0,T;V)x^{k}\in W^{1,2}(0,T;\mathbb{R}^{V}) be the solution to (P)ak,hk,x0k{}_{a^{k},h^{k},x^{k}_{0}} (k=1,2k=1,2). Then there exists a constant γ\gamma depending only on pp and G=(V,E,w)G=(V,E,w) such that

(16) sup0tT(i=1n|xi1(t)xi2(t)|2)1/2(i=1n|x0i1x0i2|2)1/2+0Th1(t)h2(t)𝑑t+γ(sup0tTx1(t)p12+sup0tTx2(t)p12)(0Ta1(t)a2(t)𝑑t)1/2.\begin{split}&\sup_{0\leq t\leq T}\left(\sum_{i=1}^{n}|x^{1}_{i}(t)-x^{2}_{i}(t)|^{2}\right)^{1/2}\leq\left(\sum_{i=1}^{n}|x^{1}_{0i}-x^{2}_{0i}|^{2}\right)^{1/2}+\int_{0}^{T}\|h^{1}(t)-h^{2}(t)\|dt\\ &\hskip 56.9055pt+\gamma\left(\sup_{0\leq t\leq T}\|x^{1}(t)\|^{\frac{p-1}{2}}+\sup_{0\leq t\leq T}\|x^{2}(t)\|^{\frac{p-1}{2}}\right)\left(\int_{0}^{T}\|a^{1}(t)-a^{2}(t)\|dt\right)^{1/2}.\end{split}
Proof..

Let ξkIKak()(xk)\xi^{k}\in\partial I_{K_{a^{k}}(\cdot)}(x^{k}) and ηkφ(xk)\eta^{k}\in\partial\varphi(x^{k}) be the sections of (P)ak,hk,x0k{}_{a^{k},h^{k},x^{k}_{0}} (k=1,2k=1,2), namely, satisfy

(xk(t))+ηk(t)+ξk(t)=hk(t)\left(x^{k}(t)\right)^{\prime}+\eta^{k}(t)+\xi^{k}(t)=h^{k}(t)

for a.e. t[0,T]t\in[0,T]. Multiply the difference of equations, which is equivalent to

(x1(t)a1(t)x2(t)+a2(t))+η1(t)η2(t)+ξ1(t)ξ2(t)\displaystyle\left(x^{1}(t)-a^{1}(t)-x^{2}(t)+a^{2}(t)\right)^{\prime}+\eta^{1}(t)-\eta^{2}(t)+\xi^{1}(t)-\xi^{2}(t)
=h1(t)h2(t)(a1(t)a2(t)),\displaystyle=h^{1}(t)-h^{2}(t)-\left(a^{1}(t)-a^{2}(t)\right)^{\prime},

by

x1(t)a1(t)x2(t)+a2(t)=(x11(t)x12(t),,xn1(t)xn2(t),0,,0)x^{1}(t)-a^{1}(t)-x^{2}(t)+a^{2}(t)=(x^{1}_{1}(t)-x^{2}_{1}(t),\ldots,x^{1}_{n}(t)-x^{2}_{n}(t),0,\ldots,0)

(recall xk(t)Kak(t)x^{k}(t)\in K_{a^{k}}(t)). According to Lemma 2.1, ξk(t)IKak(t)(xk(t))\xi^{k}(t)\in\partial I_{K_{a^{k}}(t)}(x^{k}(t)) can be represented as ξk(t)=(0,,0,ξn+1k(t),,ξn+mk(t))\xi^{k}(t)=(0,\ldots,0,\xi^{k}_{n+1}(t),\ldots,\xi^{k}_{n+m}(t)) and then fulfill

(ξ1(t)ξ2(t))(x1(t)a1(t)x2(t)+a2(t))\displaystyle\left(\xi^{1}(t)-\xi^{2}(t)\right)\cdot\left(x^{1}(t)-a^{1}(t)-x^{2}(t)+a^{2}(t)\right)
=i=1n0(xi1(t)xi2(t))+j=1m(ξn+j1(t)ξn+j2(t))0=0.\displaystyle=\sum_{i=1}^{n}0\cdot(x^{1}_{i}(t)-x^{2}_{i}(t))+\sum_{j=1}^{m}(\xi^{1}_{n+j}(t)-\xi^{2}_{n+j}(t))\cdot 0=0.

For the same reason, (a1(t)a2(t))(x1(t)a1(t)x2(t)+a2(t))=0\left(a^{1}(t)-a^{2}(t)\right)^{\prime}\cdot\left(x^{1}(t)-a^{1}(t)-x^{2}(t)+a^{2}(t)\right)=0 also holds by (ak(t))=(0,,0,(a1k(t)),,(amk(t)))(a^{k}(t))^{\prime}=(0,\ldots,0,(a^{k}_{1}(t))^{\prime},\ldots,(a^{k}_{m}(t))^{\prime}). Therefore we can derive

(17) 12ddti=1n|xi1(t)xi2(t)|2(η1(t)+η2(t))a1(t)a2(t)+h1(t)h2(t)(i=1n|xi1(t)xi2(t)|2)1/2,\begin{split}\frac{1}{2}\frac{d}{dt}\sum_{i=1}^{n}|x^{1}_{i}(t)-x^{2}_{i}(t)|^{2}&\leq\left(\|\eta^{1}(t)\|+\|\eta^{2}(t)\|\right)\|a^{1}(t)-a^{2}(t)\|\\ &\hskip 28.45274pt+\|h^{1}(t)-h^{2}(t)\|\left(\sum_{i=1}^{n}|x^{1}_{i}(t)-x^{2}_{i}(t)|^{2}\right)^{1/2},\end{split}

which provides (16) by Lemma 2.3 with

Ξ(t)=(i=1n|xi1(t)xi2(t)|2)1/2,g(t)=h1(t)h2(t),\displaystyle\Xi(t)=\left(\sum_{i=1}^{n}|x^{1}_{i}(t)-x^{2}_{i}(t)|^{2}\right)^{1/2},~{}~{}~{}~{}~{}g(t)=\|h^{1}(t)-h^{2}(t)\|,
κ=(i=1n|x0i1x0i2|2+20T(η1(t)+η2(t))a1(t)a2(t)𝑑t)1/2\displaystyle\kappa=\left(\sum_{i=1}^{n}|x^{1}_{0i}-x^{2}_{0i}|^{2}+2\int_{0}^{T}\left(\|\eta^{1}(t)\|+\|\eta^{2}(t)\|\right)\|a^{1}(t)-a^{2}(t)\|dt\right)^{1/2}

and (12). ∎

Remark 4.2.

As shown in [7, Remark 2.7], the hypergraph Laplacian is not strongly monotone, i.e., it may be φ(x1)=φ(x2)\partial\varphi(x^{1})=\partial\varphi(x^{2}) but x1x2x^{1}\neq x^{2}. Hence it is not easy to deduce better information from (η1(t)η2(t))(x1(t)x2(t))(\eta^{1}(t)-\eta^{2}(t))\cdot(x^{1}(t)-x^{2}(t)) in (17) than above in general.

Remark 4.3.

Theorem 4.1 implies that the solution to (P)a,h,x0{}_{a,h,x_{0}} is 1/21/2-Hölder continuous with respect to the given data a(t)a(t). This seems to be optimal and it is difficult to derive better estimate in general. Indeed, let E={V}E=\{V\}, xk(t)=(x1k(t),,xnk(t),a1k(t),,amk(t))x^{k}(t)=(x^{k}_{1}(t),\ldots,x^{k}_{n}(t),a^{k}_{1}(t),\ldots,a^{k}_{m}(t)) be a solution to (P)ak,hk,x0k{}_{a^{k},h^{k},x^{k}_{0}} (k=1,2k=1,2), and ηk(t)φ(xk(t))\eta^{k}(t)\in\partial\varphi(x^{k}(t)) be the section of (P)ak,hk,x0k{}_{a^{k},h^{k},x^{k}_{0}}. Moreover, we assume the initial data for x1x^{1} satisfies

xi11(0)<aj1(0)<xi21(0)j=1,,mx^{1}_{i_{1}}(0)<a^{1}_{j}(0)<x^{1}_{i_{2}}(0)~{}~{}~{}~{}\forall j=1,\ldots,m

with some i1,i2{1,,n}i_{1},i_{2}\in\{1,\ldots,n\} and for x2x^{2} fulfills

aj12(0)<xi2(0)<aj22(0)i=1,,na^{2}_{j_{1}}(0)<x^{2}_{i}(0)<a^{2}_{j_{2}}(0)~{}~{}~{}~{}\forall i=1,\ldots,n

with some j1,j2{1,,m}j_{1},j_{2}\in\{1,\ldots,m\}. By the continuity of solutions and aka^{k}, there exists some T0>0T_{0}>0 such that the solutions still satisfy xi11(t)<aj1(t)<xi21(t)x^{1}_{i_{1}}(t)<a^{1}_{j}(t)<x^{1}_{i_{2}}(t) and aj12(t)<xi2(t)<aj22(t)a^{2}_{j_{1}}(t)<x^{2}_{i}(t)<a^{2}_{j_{2}}(t) (i=1,,n\forall i=1,\ldots,n, j=1,,m\forall j=1,\ldots,m) on [0,T0][0,T_{0}]. Then we can see by (2) that ηk(t)φ(xk(t))\eta^{k}(t)\in\partial\varphi(x^{k}(t)) with t[0,T0]t\in[0,T_{0}] can be represented by

η1(t)=(η11(t),,ηn1(t),0,,0),η2(t)=(0,,0,ηn+12(t),,ηn+m2(t)).\eta^{1}(t)=(\eta^{1}_{1}(t),\ldots,\eta^{1}_{n}(t),0,\ldots,0),~{}~{}~{}~{}\eta^{2}(t)=(0,\ldots,0,\eta^{2}_{n+1}(t),\ldots,\eta^{2}_{n+m}(t)).

Hence we can obtain

(η1(t)η2(t))(x1(t)a1(t)x2(t)+a2(t))\displaystyle(\eta^{1}(t)-\eta^{2}(t))\cdot(x^{1}(t)-a^{1}(t)-x^{2}(t)+a^{2}(t))
=η1(t)(x1(t)a1(t)x2(t)+a2(t))=i=1nηi1(t)(xi1(t)xi2(t)),\displaystyle=\eta^{1}(t)\cdot(x^{1}(t)-a^{1}(t)-x^{2}(t)+a^{2}(t))=\sum_{i=1}^{n}\eta^{1}_{i}(t)(x^{1}_{i}(t)-x^{2}_{i}(t)),

which appears in the estimate above. It is not obvious whether we can deduce the term x1(t)x2(t)α\|x^{1}(t)-x^{2}(t)\|^{\alpha} (α>1)(\alpha>1) from this.

5 Asymptotic Behavior of Global Solution

Throughout this section, let x(t)x(t) stand for a unique solution to (P)a,h,x0{}_{a,h,x_{0}}. According to Theorem 3.1, the solution can be globally extended if

(18) aWloc1,2(0,;V),hLloc2(0,;V).a\in W^{1,2}_{\text{loc}}(0,\infty;\mathbb{R}^{V}),~{}~{}~{}~{}h\in L^{2}_{\text{loc}}(0,\infty;\mathbb{R}^{V}).

We discuss the global behavior of solution. In addition to the above, we assume

(19) a,hL1(0,;V)a^{\prime},h^{\prime}\in L^{1}(0,\infty;\mathbb{R}^{V})

so that a=limta(t)a_{\infty}=\lim_{t\to\infty}a(t) and h=limth(t)h_{\infty}=\lim_{t\to\infty}h(t) can be determined uniquely.

We first consider the case where aa is independent of the time. Henceforth, Ka(t)K_{a}(t) is abbreviated to KaK_{a} when aa is a constant vector.

Proposition 5.1.

Let a=(0,,0,a1,,am)Va=(0,\ldots,0,a_{1},\ldots,a_{m})\in\mathbb{R}^{V} be a constant vector and hh satisfy (18) and (19). If hR(φ+IKa)h_{\infty}\in R(\partial\varphi+\partial I_{K_{a}}) and hhL1(0,;V)h-h_{\infty}\in L^{1}(0,\infty;\mathbb{R}^{V}), then x=limtx(t)x_{\infty}=\lim_{t\to\infty}x(t) is determined uniquely and satisfies φ(x)+IKa(x)h\partial\varphi(x_{\infty})+\partial I_{K_{a}}(x_{\infty})\ni h_{\infty}, that is, xx_{\infty} is a stable solution to (P)a,h,x{}_{a,h_{\infty},x_{\infty}}.

Proof..

We can apply [1, Théorème 3.11], since (φ+IKa)=φ+IKa\partial(\varphi+I_{K_{a}})=\partial\varphi+\partial I_{K_{a}} is time-independent. ∎

Remark 5.2.

Let a=(0,,0,a1,,am)a=(0,\ldots,0,a_{1},\ldots,a_{m}) be a constant vector. By the definition of subdifferential (9), hR(φ+IKa)h_{\infty}\in R(\partial\varphi+\partial I_{K_{a}}) holds if and only if the closed convex set {xV;Φ(x)=minyVΦ(y)}\{x\in\mathbb{R}^{V};~{}\Phi(x)=\min_{y\in\mathbb{R}^{V}}\Phi(y)\} is not empty, where Φ:V\Phi:\mathbb{R}^{V}\to\mathbb{R} is a proper lower semi-continuous convex function defined by

Φ(x):={φ(x)+IKa(x)hx if xKa,+ if xKa.\Phi(x):=\begin{cases}~{}~{}\displaystyle\varphi(x)+I_{K_{a}}(x)-h_{\infty}\cdot x~{}~{}&~{}~{}\text{ if }x\in K_{a},\\ ~{}~{}\displaystyle+\infty~{}~{}&~{}~{}\text{ if }x\not\in K_{a}.\end{cases}

From this and φ,IKa0\varphi,I_{K_{a}}\geq 0, we can derive 0R(φ+IKa)0\in R(\partial\varphi+\partial I_{K_{a}}) for every p1p\geq 1. Since Theorem 2.2 implies

Φ(x)12pCpxp(n+1)pCp(j=1m|aj|)phx,\Phi(x)\geq\frac{1}{2^{p}C^{p}}\|x\|^{p}-\frac{(n+1)^{p}}{C^{p}}\left(\sum_{j=1}^{m}|a_{j}|\right)^{p}-h_{\infty}\cdot x,

there exists a global minimizer for every hh_{\infty}, i.e., R(φ+IKa)=VR(\partial\varphi+\partial I_{K_{a}})=\mathbb{R}^{V} holds if p>1p>1. When p=1p=1, however, we can see that R(φ+IKa)VR(\partial\varphi+\partial I_{K_{a}})\neq\mathbb{R}^{V}. Indeed,

a0,h=(4(#E)maxeEw(e),0,,0),xμ=(μ,0,,0)Ka(μ=1,2,)a\equiv 0,~{}~{}~{}h_{\infty}=(4({\#}E)\max_{e\in E}w(e),0,\ldots,0),~{}~{}~{}x^{\mu}=(\mu,0,\ldots,0)\in K_{a}~{}~{}~{}(\mu=1,2,\ldots)

satisfy Φ(xμ)\Phi(x_{\mu})\to-\infty as μ\mu\to\infty (see (12)).

We next deal with the case where aa depends on tt and h(t),a(t)0h(t),a(t)\to 0 as tt\to\infty.

Theorem 5.3.

Let (18), (19), and hL1(0,;V)L2(0,;V)h\in L^{1}(0,\infty;\mathbb{R}^{V})\cap L^{2}(0,\infty;\mathbb{R}^{V}) be assumed. If aLp(0,;V)a\in L^{p}(0,\infty;\mathbb{R}^{V}) and aL2(0,;V)a^{\prime}\in L^{2}(0,\infty;\mathbb{R}^{V}) are satisfied, then x(t)0x(t)\to 0 as tt\to\infty. In particular, if h=a0h=a\equiv 0, then there exists some constant γ>0\gamma>0 depending only on pp and G=(V,E,w)G=(V,E,w) such that

(20) x(t){(x02pγt)+12p if 1p<2,x0exp(γt) if p=2,(x0(p2)+γt)1p2 if p>2,\|x(t)\|\leq\begin{cases}~{}~{}\displaystyle\left(\|x_{0}\|^{2-p}-\gamma t\right)^{\frac{1}{2-p}}_{+}~{}~{}&~{}~{}\text{ if }1\leq p<2,\\ ~{}~{}\displaystyle\|x_{0}\|\exp\left(-\gamma t\right)~{}~{}&~{}~{}\text{ if }p=2,\\ ~{}~{}\displaystyle\left(\|x_{0}\|^{-(p-2)}+\gamma t\right)^{-\frac{1}{p-2}}~{}~{}&~{}~{}\text{ if }p>2,\end{cases}

where (s)+:=max{0,s}(s)_{+}:=\max\{0,s\} stands for the positive part of ss.

Proof..

Henceforth, let c0,C0c_{0},C_{0} denote general constants independent of TT. Note that the assumptions lead to a,hL(0,;V)a,h\in L^{\infty}(0,\infty;\mathbb{R}^{V}) and a=h=0a_{\infty}=h_{\infty}=0.

Testing (P)a,h,x0{}_{a,h,x_{0}}, which is equivalent to

(x(t)a(t))+φ(x(t))+IKa(t)(x(t))h(t)a(t),(x(t)-a(t))^{\prime}+\partial\varphi(x(t))+\partial I_{K_{a}(t)}(x(t))\ni h(t)-a^{\prime}(t),

by (xa)=(x1(t),,xn(t),0,,0)(x-a)=(x_{1}(t),\ldots,x_{n}(t),0,\ldots,0) (recall (6)), we obtain

(21) 12ddti=1n|xi(t)|2+φ(x(t))φ(a(t))h(t)(i=1n|xi(t)|2)1/2.\frac{1}{2}\frac{d}{dt}\sum_{i=1}^{n}|x_{i}(t)|^{2}+\varphi(x(t))-\varphi(a(t))\leq\|h(t)\|\left(\sum_{i=1}^{n}|x_{i}(t)|^{2}\right)^{1/2}.

Here we use the fact that ξ(t)(x(t)a(t))=0\xi(t)\cdot(x(t)-a(t))=0 holds for every ξ(t)IKa(t)(x(t))\xi(t)\in\partial I_{K_{a}(t)}(x(t)) by Lemma 2.1 and a(t)(x(t)a(t))=0a^{\prime}(t)\cdot(x(t)-a(t))=0 by a(t)=(0,,0,a1(t),,am(t))a^{\prime}(t)=(0,\ldots,0,a^{\prime}_{1}(t),\ldots,a^{\prime}_{m}(t)). By Lemma 2.3, we have for arbitrary T>0T>0

(22) sup0tT(i=1n|xi(t)|2)1/2(i=1n|x0i|2+20Tφ(a(t))𝑑t)1/2+0Th(t)𝑑t.\sup_{0\leq t\leq T}\left(\sum_{i=1}^{n}|x_{i}(t)|^{2}\right)^{1/2}\leq\left(\sum_{i=1}^{n}|x_{0i}|^{2}+2\int_{0}^{T}\varphi(a(t))dt\right)^{1/2}+\int_{0}^{T}\|h(t)\|dt.

Since aLp(0,;V)a\in L^{p}(0,\infty;\mathbb{R}^{V}) is assumed and φ(a(t))C0a(t)p\varphi(a(t))\leq C_{0}\|a(t)\|^{p} holds by (12), it follows that sup0t<x(t)\sup_{0\leq t<\infty}\|x(t)\| is bounded. This together with (12) implies that the section ηφ(x)\eta\in\partial\varphi(x) of (P)a,h,x0{}_{a,h,x_{0}} satisfies sup0t<η(t)C0\sup_{0\leq t<\infty}\|\eta(t)\|\leq C_{0}. Moreover, Theorem 2.2 yields

φ(x(t))12pC(i=1n|xi(t)|)pnp2pC(j=1m|aj(t)|)p.\varphi(x(t))\geq\frac{1}{2^{p}C}\left(\sum_{i=1}^{n}|x_{i}(t)|\right)^{p}-\frac{n^{p}}{2^{p}C}\left(\sum_{j=1}^{m}|a_{j}(t)|\right)^{p}.

Hence integrating (21) over [0,)[0,\infty), we also get 0x(t)p𝑑tC0\int_{0}^{\infty}\|x(t)\|^{p}dt\leq C_{0}.

Next multiplying (P)a,h,x0{}_{a,h,x_{0}} by (xa)=(x1(t),,xn(t),0,,0)(x-a)^{\prime}=(x^{\prime}_{1}(t),\ldots,x^{\prime}_{n}(t),0,\ldots,0) and using the boundedness of η(t)\eta(t), we have

12i=1n|xi(t)|2+ddtφ(x(t))C0a(t)+12h(t)2.\frac{1}{2}\sum_{i=1}^{n}|x^{\prime}_{i}(t)|^{2}+\frac{d}{dt}\varphi(x(t))\leq C_{0}\|a^{\prime}(t)\|+\frac{1}{2}\|h(t)\|^{2}.

We here used Lemma 2.1 and a(t)(x(t)a(t))=0a^{\prime}(t)\cdot(x(t)-a(t))^{\prime}=0. From aL1(0,;V)a^{\prime}\in L^{1}(0,\infty;\mathbb{R}^{V}) and hL2(0,;V)h\in L^{2}(0,\infty;\mathbb{R}^{V}), we can derive i=1n0|xi(t)|2𝑑tC0\sum_{i=1}^{n}\int_{0}^{\infty}|x^{\prime}_{i}(t)|^{2}dt\leq C_{0}.

Define the ω\omega-limit set of the solution xx to (P)a,h,x0{}_{a,h,x_{0}} by

ω(a,h,x0):={yV;{tk}k s.t. tkx(tk)y as k}.\omega(a,h,x_{0}):=\{y\in\mathbb{R}^{V};~{}\exists\{t_{k}\}_{k\in\mathbb{N}}~{}~{}\text{ s.t. }~{}~{}t_{k}\to\infty~{}~{}x(t_{k})\to y~{}~{}\text{ as }k\to\infty\}.

By the global boundedness of x(t)x(t) given above, ω(a,h,x0)\omega(a,h,x_{0})\neq\varnothing holds under the assumptions of Theorem 5.3. Let xω(a,h,x0)x_{\infty}\in\omega(a,h,x_{0}) and {tk}k\{t_{k}\}_{k\in\mathbb{N}} satisfy x(tk)xx(t_{k})\to x_{\infty} and tkt_{k}\to\infty as kk\to\infty. Note that xKax_{\infty}\in K_{a_{\infty}} since a(t)aa(t)\to a_{\infty} and xn+j(t)=aj(t)x_{n+j}(t)=a_{j}(t).

From xL2(0,;V)x^{\prime}\in L^{2}(0,\infty;\mathbb{R}^{V}), it follows that for every s[0,1]s\in[0,1]

(i=1n|xi(tk+s)xi(tk)|2)1/2(i=1ntktk+1|xi(τ)|2𝑑τ)1/20\left(\sum_{i=1}^{n}|x_{i}(t_{k}+s)-x_{i}(t_{k})|^{2}\right)^{1/2}\leq\left(\sum_{i=1}^{n}\int_{t_{k}}^{t_{k}+1}|x^{\prime}_{i}(\tau)|^{2}d\tau\right)^{1/2}\to 0

as tkt_{k}\to\infty, that is, xi(tk+s)xix_{i}(t_{k}+s)\to x_{\infty i} holds uniformly with respect to s[0,1]s\in[0,1].

We here define XkW1,2(0,1;V)X_{k}\in W^{1,2}(0,1;\mathbb{R}^{V}) by Xk(s):=x(tk+s)X_{k}(s):=x(t_{k}+s). Then it is easy to see that Xk(s)xX_{k}(s)\to x_{\infty} uniformly in [0,1][0,1] and XkX_{k} is a solution to

Xk(s)+φ(Xk(s))+IKa(tk+s)(Xk(s))h(tk+s).X^{\prime}_{k}(s)+\partial\varphi(X_{k}(s))+\partial I_{K_{a}(t_{k}+s)}(X_{k}(s))\ni h(t_{k}+s).

Let ζk(s)φ(Xk(s))+IKa(tk+s)(Xk(s))\zeta_{k}(s)\in\partial\varphi(X_{k}(s))+\partial I_{K_{a}(t_{k}+s)}(X_{k}(s)) be the section of this inclusion, then

(01ζk(s)2𝑑s)1/2\displaystyle\left(\int_{0}^{1}\|\zeta_{k}(s)\|^{2}ds\right)^{1/2} (01Xk(s)2𝑑s)1/2+(01h(tk+s)2𝑑s)1/2\displaystyle\leq\left(\int_{0}^{1}\|X^{\prime}_{k}(s)\|^{2}ds\right)^{1/2}+\left(\int_{0}^{1}\|h(t_{k}+s)\|^{2}ds\right)^{1/2}
=(tktk+1x(s)2𝑑s)1/2+(tktk+1h(s)2𝑑s)1/2.\displaystyle=\left(\int_{t_{k}}^{t_{k}+1}\|x^{\prime}(s)\|^{2}ds\right)^{1/2}+\left(\int_{t_{k}}^{t_{k}+1}\|h(s)\|^{2}ds\right)^{1/2}.

Hence ζk0\zeta_{k}\to 0 in L2(0,1;V)L^{2}(0,1;\mathbb{R}^{V}) by a,hL2(0,;V)a^{\prime},h\in L^{2}(0,\infty;\mathbb{R}^{V}).

Choose y1,,yny_{1},\ldots,y_{n}\in\mathbb{R} arbitrarily and define

Yk(s):=(y1,,yn,a1(tk+s),,am(tk+s))Ka(tk+s).Y_{k}(s):=(y_{1},\ldots,y_{n},a_{1}(t_{k}+s),\ldots,a_{m}(t_{k}+s))\in K_{a}(t_{k}+s).

Obviously, Yk(s)Y_{k}(s) converges to y:=(y1,,yn,y_{\infty}:=(y_{1},\ldots,y_{n}, a1,,am)Kaa_{\infty 1},\ldots,a_{\infty m})\in K_{a_{\infty}} as kk\to\infty uniformly with respect to s[0,1]s\in[0,1]. By the definition of subdifferential (9) and (φ+IKa(tk+s))=φ+IKa(tk+s)\partial(\varphi+I_{K_{a}(t_{k}+s)})=\partial\varphi+\partial I_{K_{a}(t_{k}+s)}, we obtain

01ζk(s)(Yk(s)Xk(s))𝑑s01(φ(Yk(s))φ(Xk(s)))𝑑s.\int_{0}^{1}\zeta_{k}(s)\cdot(Y_{k}(s)-X_{k}(s))ds\leq\int_{0}^{1}\left(\varphi(Y_{k}(s))-\varphi(X_{k}(s))\right)ds.

Since a,xL(0,;V)a,x\in L^{\infty}(0,\infty;\mathbb{R}^{V}) and (12), the dominated convergence theorem is applicable to this and

001(φ(y)φ(x))𝑑s=(φ(y)+IKa(y))(φ(x)+IKa(x))0\leq\int_{0}^{1}\left(\varphi(y_{\infty})-\varphi(x_{\infty})\right)ds=\left(\varphi(y_{\infty})+I_{K_{a_{\infty}}}(y_{\infty})\right)-\left(\varphi(x_{\infty})+I_{K_{a_{\infty}}}(x_{\infty})\right)

holds for every y1,,yny_{1},\ldots,y_{n}\in\mathbb{R}, which implies that φ+IKa\varphi+I_{K_{a_{\infty}}} attains its minimum at xx_{\infty}. When a=0a_{\infty}=0, the minimum of φ+IKa\varphi+I_{K_{a_{\infty}}} coincides with 0. From this fact and (11), we can derive x=(0,,0)x_{\infty}=(0,\ldots,0). Since the limit is determined independently of the choice of subsequences {tk}\{t_{k}\}, the whole trajectory {x(t)}\{x(t)\} also tends to 0. Therefore we can conclude that ω(a,0,x0)={0}\omega(a,0,x_{0})=\{0\}.

Especially, if a,h0a,h\equiv 0, then (21) becomes

12ddti=1n|xi(t)|2+φ(x(t))φ(0)=0.\frac{1}{2}\frac{d}{dt}\sum_{i=1}^{n}|x_{i}(t)|^{2}+\varphi(x(t))\leq\varphi(0)=0.

Furthermore, (10) leads to Cφ(x(t))(i=1n|xi(t)|)pC\varphi(x(t))\geq\left(\sum_{i=1}^{n}|x_{i}(t)|\right)^{p}. Hence multiplying (P)0,0,x0{}_{0,0,x_{0}} by x=(xa)x=(x-a), we get

12ddtx(t)2+c0x(t)p0\frac{1}{2}\frac{d}{dt}\|x(t)\|^{2}+c_{0}\|x(t)\|^{p}\leq 0

which immediately yields (20). ∎

We next consider that case where a0a_{\infty}\neq 0. When 0φ(a(t))𝑑t<\int_{0}^{\infty}\varphi(a(t))dt<\infty, there exists some subsequence {tk}k\{t_{k}\}_{k\in\mathbb{N}} such that tkt_{k}\to\infty and φ(a(tk))φ(a)=0\varphi(a(t_{k}))\to\varphi(a_{\infty})=0, which together with (11) leads to a=0a_{\infty}=0. This implies that sup0t<x(t)<\sup_{0\leq t<\infty}\|x(t)\|<\infty can not be deduced via a priori estimate (22). We here give another way to establish uniform estimates by using Theorem 4.1.

Theorem 5.4.

Let 1p<31\leq p<3, (18), and (19) be satisfied. Moreover, suppose that

aL2(0,;V),hR(φ+IKa),\displaystyle a^{\prime}\in L^{2}(0,\infty;\mathbb{R}^{V}),~{}~{}~{}h_{\infty}\in R(\partial\varphi+\partial I_{K_{a_{\infty}}}),
aaL1(0,;V),hhL1(0,;V)L2(0,;V).\displaystyle a-a_{\infty}\in L^{1}(0,\infty;\mathbb{R}^{V}),~{}~{}~{}h-h_{\infty}\in L^{1}(0,\infty;\mathbb{R}^{V})\cap L^{2}(0,\infty;\mathbb{R}^{V}).

Then x=limtx(t)x_{\infty}=\lim_{t\to\infty}x(t) is determined uniquely and satisfies φ(x)+IKa(x)h\partial\varphi(x_{\infty})+\partial I_{K_{a_{\infty}}}(x_{\infty})\ni h_{\infty}.

Proof..

Let zVz_{\infty}\in\mathbb{R}^{V} satisfy hφ(z)+IKa(z)h_{\infty}\in\partial\varphi(z_{\infty})+\partial I_{K_{a_{\infty}}}(z_{\infty}), which means that zz_{\infty} is a solution to (P)a,h,z{}_{a_{\infty},h_{\infty},z_{\infty}}. By Theorem 4.1 with (a2,h2,x02)=(a,h,z)(a^{2},h^{2},x^{2}_{0})=(a_{\infty},h_{\infty},z_{\infty}), we obtain

sup0tT(i=1n|xi(t)zi|2)1/2C0(sup0tTx(t)p12+1),\sup_{0\leq t\leq T}\left(\sum_{i=1}^{n}|x_{i}(t)-z_{\infty i}|^{2}\right)^{1/2}\leq C_{0}\left(\sup_{0\leq t\leq T}\|x(t)\|^{\frac{p-1}{2}}+1\right),

where C0C_{0} is a general constant independent of T>0T>0. Then when p=1p=1, we can obtain sup0t<x(t)<\sup_{0\leq t<\infty}\|x(t)\|<\infty. Otherwise, let Z(T):=sup0tT(i=1n|xi(t)zi|2)1/2Z(T):=\sup_{0\leq t\leq T}\left(\sum_{i=1}^{n}|x_{i}(t)-z_{\infty i}|^{2}\right)^{1/2}, which satisfies Z(T)C0(Z(T)p12+1)Z(T)\leq C_{0}(Z(T)^{\frac{p-1}{2}}+1) from above. Hence if 1<p<31<p<3, we obtain Z(T)max{(C0+1)2/(3p),C02/(p1)}Z(T)\leq\max\{(C_{0}+1)^{2/(3-p)},C_{0}^{2/(p-1)}\}. By the arbitrariness of TT, it follows that sup0t<x(t)<\sup_{0\leq t<\infty}\|x(t)\|<\infty, which leads to sup0t<η(t)<\sup_{0\leq t<\infty}\|\eta(t)\|<\infty. Testing (P)a,h,x0{}_{a,h,x_{0}} by (xa)(x-a)^{\prime}, we have

x(t)a(t)2+ddtφ(x(t))\displaystyle\|x^{\prime}(t)-a^{\prime}(t)\|^{2}+\frac{d}{dt}\varphi(x(t))
η(t)a(t)+h(t)hx(t)a(t)+h(x(t)a(t))\displaystyle\leq\|\eta(t)\|\|a^{\prime}(t)\|+\|h(t)-h_{\infty}\|\|x^{\prime}(t)-a^{\prime}(t)\|+h_{\infty}\cdot(x(t)-a(t))^{\prime}

Here 0Th(x(t)a(t))𝑑t=h(x(T)a(T)x0+a(0))\int_{0}^{T}h_{\infty}\cdot(x(t)-a(t))^{\prime}dt=h_{\infty}\cdot(x(T)-a(T)-x_{0}+a(0)) is uniformly bounded by (19) and sup0t<x(t)<\sup_{0\leq t<\infty}\|x(t)\|<\infty. Hence 0x(t)2𝑑t<\int_{0}^{\infty}\|x^{\prime}(t)\|^{2}dt<\infty also holds.

Fix xω(a,h,x0)x_{\infty}\in\omega(a,h,x_{0}) and let {tk}k\{t_{k}\}_{k\in\mathbb{N}} satisfy tkt_{k}\to\infty and x(tk)xx(t_{k})\to x_{\infty} as kk\to\infty. As above, define Xk:=x(tk+)W1,2(0,1;V)X_{k}:=x(t_{k}+\cdot)\in W^{1,2}(0,1;\mathbb{R}^{V}), which is the solution to

Xk(s)+φ(Xk(s))+IKa(tk+s)(Xk(s))hh(tk+s)h.X^{\prime}_{k}(s)+\partial\varphi(X_{k}(s))+\partial I_{K_{a}(t_{k}+s)}(X_{k}(s))-h_{\infty}\ni h(t_{k}+s)-h_{\infty}.

Then the section ζk(s)φ(Xk(s))+IKa(tk+s)(Xk(s))\zeta_{k}(s)\in\partial\varphi(X_{k}(s))+\partial I_{K_{a}(t_{k}+s)}(X_{k}(s)) of this inclusion satisfies

(01ζk(s)h2𝑑s)1/2=(tktk+1x(s)h(s)+h2𝑑s)1/20,\left(\int_{0}^{1}\|\zeta_{k}(s)-h_{\infty}\|^{2}ds\right)^{1/2}=\left(\int_{t_{k}}^{t_{k}+1}\|x^{\prime}(s)-h(s)+h_{\infty}\|^{2}ds\right)^{1/2}\to 0,

namely, ζkh\zeta_{k}\to h_{\infty} in L2(0,1;V)L^{2}(0,1;\mathbb{R}^{V}). For every y1,,yny_{1},\ldots,y_{n}\in\mathbb{R} we can obtain

h(yx)(φ(y)+IKa(y))(φ(x)+IKa(x)),h_{\infty}\cdot(y_{\infty}-x_{\infty})\leq\left(\varphi(y_{\infty})+I_{K_{a_{\infty}}}(y_{\infty})\right)-\left(\varphi(x_{\infty})+I_{K_{a_{\infty}}}(x_{\infty})\right),

where y=(y1,,yn,a1,,am)y_{\infty}=(y_{1},\ldots,y_{n},a_{\infty 1},\ldots,a_{\infty m}). Therefore yφ(y)+IKa(y)hyy\mapsto\varphi(y)+I_{K_{a_{\infty}}}(y)-h_{\infty}\cdot y attains its minimum at xx_{\infty}, which implies that φ(x)+IKa(x)h\partial\varphi(x_{\infty})+\partial I_{K_{a_{\infty}}}(x_{\infty})\ni h_{\infty}.

Again, choose xω(a,h,x0)x_{\infty}\in\omega(a,h,x_{0}) arbitrarily and let {tk}k\{t_{k}\}_{k\in\mathbb{N}} satisfy x(tk)xx(t_{k})\to x_{\infty}. Since xx_{\infty} is a solution to (P)a,h,x{}_{a_{\infty},h_{\infty},x_{\infty}}, we can derive from Theorem 4.1 with t=0t=0 replaced by t=tkt=t_{k}

(i=1n|xi(t+tk)xi|2)1/2(i=1n|xi(tk)xi|2)1/2\displaystyle\left(\sum_{i=1}^{n}|x_{i}(t+t_{k})-x_{\infty i}|^{2}\right)^{1/2}\leq\left(\sum_{i=1}^{n}|x_{i}(t_{k})-x_{\infty i}|^{2}\right)^{1/2}
+C0(tka(t)a𝑑t)1/2+tkh(t)h𝑑t\displaystyle\hskip 85.35826pt+C_{0}\left(\int_{t_{k}}^{\infty}\|a(t)-a_{\infty}\|dt\right)^{1/2}+\int_{t_{k}}^{\infty}\|h(t)-h_{\infty}\|dt

for every t>0t>0. Hence x(t)xx(t)\to x_{\infty} can be assured regardless of the choice of subsequences, that is, the limit xx_{\infty} is determined uniquely. ∎

Remark 5.5.

Under the same assumptions for aa and hh as in Theorem 5.4, we can use [5, Theorem 2] and show the uniform boundedness of x(t)\|x(t)\| without any restriction for p[1,)p\in[1,\infty) (see also [8, Theorem 2.2] and [14, Theorem 2.2] for global boundedness of solution). Hence we can derive Theorem 5.4 for every p[1,)p\in[1,\infty) by referring known abstract results. We here mention that, however, the convergence of whole sequence {x(t)}\{x(t)\} without extraction specific subsequence can be obtained in our result thanks to the continuous dependence of solutions (Theorem 4.1), which can not be assured only by [5, 8, 14]. Moreover, we recall that the decay estimate (20) in Theorem 5.3 is derived from the Poincaré type inequality (Theorem 2.2), that is, the nature of the hypergraph Laplacian.

Acknowedgements

T. Fukao is supported by JSPS Grant-in-Aid for Scientific Research (C) (No.21K03309). M. Ikeda is supported by JST CREST Grant (No.JPMJCR1913) and JSPS Grant-in-Aid for Young Scientists Research (No.19K14581). S. Uchida is supported by JSPS Fund for the Promotion of Joint International Research (Fostering Joint International Research (B)) (No.18KK0073) and Sumitomo Foundation Fiscal 2022 Grant for Basic Science Research Projects (No.2200250).

References

  • [1] H. Brézis, Opérateurs maximaux monotones et semi-groupes de contractions dans les espaces de Hilbert, North-Holland, Amsterdam, 1973.
  • [2] C. Combari; M. Laghdir; L. Thibault, A note on subdifferentials of convex composite functionals, Arch. Math. 67 (1996), 239–252.
  • [3] K. Fujii; T. Soma; Y. Yoshida, Polynomial-Time Algorithms for Submodular Laplacian Systems, Theoret. Comput. Sci. 892 (2021), 170–186.
  • [4] T. Fukao; N. Kenmochi, Parabolic variational inequalities with weakly time-dependent constraints, Adv. Math. Sci. Appl. 23(2) (2013), 365–395.
  • [5] H. Furuya; K. Miyashiba; N. Kenmochi, Asymptotic behavior of solutions to a class of nonlinear evolution equations, J. Differential Equations 62(1) (1986), 73–94.
  • [6] M. Ikeda; A. Miyauchi; Y. Takai; Y. Yoshida, Finding cheeger cuts in hypergraphs via heat equation, Theoret. Comput. Sci. 930, (2022), 1–23.
  • [7] M. Ikeda; S. Uchida, Nonlinear evolution equation associated with hypergraph Laplacian, preprint (https://arxiv.org/abs/2107.14693).
  • [8] A. Ito; N. Yamazaki; N. Kenmochi, Attractors of nonlinear evolution systems generated by time-dependent subdifferentials in Hilbert spaces, 327–349, in Dynamical Systems and Differential Equations Vol. 1, Southwest Missouri State University, Springfield, 1998.
  • [9] N. Kenmochi, Solvability of nonlinear evolution equations with time-dependent constraints and applications, Bull. Fac. Edu. Chiba Univ. 30 (1981), 1–87.
  • [10] J. Kigami, A harmonic calculus on the Sierpinski spaces, Japan J. Appl. Math. 6 (1989), 259-290.
  • [11] J. Kigami, Analysis on fractals, Cambridge Tracts in Mathematics 143, Cambridge University Press, Cambridge, 2001.
  • [12] P. Li; O. Milenkovic, Submodular hypergraphs: pp-Laplacians, Cheeger inequalities and spectral clustering, Proc. 35th Internat. Conf. on Machine Learning, PMLR 80 (2018), 3014–3023.
  • [13] B. S Mordukhovich; N. M. Nam, An Easy Path to Convex Analysis and Appli- cations, Morgan and Claypool Publishers, Williston, VT, 2014.
  • [14] K. Shirakawa; A. Ito; N. Yamazaki; N. Kenmochi, Asymptotic Stability for Evolution Equations Governed by Subdifferentials, 287–310, in Recent Developments in Domain Decomposition Methods and Flow Problems, Gakuto Intern. Ser., Math. Sci. Appl. Vol. 11, Gakkotosho, Tokyo, 1998.
  • [15] Y. Takai; A. Miyauchi; M. Ikeda; Y. Yoshida, Hypergraph clustering based on PageRank, Proc. 26th ACM SIGKDD Internat. Conf. on Knowledge Discovery & Data Mining (2020), 1970–1978.
  • [16] Y. Yoshida, Cheeger Inequalities for Submodular Transformations, Proc. 2019 Annu. ACM-SIAM Symp. on Discrete Algorithms (SODA) (2019), 2582–2601.

Fukao Takeshi
Department of Mathematics,
Kyoto University of Education,
1 Fujinomori, Fukakusa,
Fushimi-ku, Kyoto,
612-8522, JAPAN.

E-mail address: fukao@kyokyo-u.ac.jp


Masahiro Ikeda
Department of Mathematics,
Faculty of Science and Technology,
Keio University,
3-14-1 Hiyoshi Kohoku-ku, Yokohama,
223-8522, JAPAN/
Center for Advanced Intelligence
Project, RIKEN, Tokyo,
103-0027, JAPAN.

E-mail address: masahiro.ikeda@keio.jp/
masahiro.ikeda@riken.jp


Shun Uchida
Department of Integrated Science and Technology,
Faculty of Science and Technology,
Oita University,
700 Dannoharu, Oita City, Oita Pref.,
870-1192, JAPAN.

E-mail address: shunuchida@oita-u.ac.jp