Next Article in Journal
Bayesian Variable Selection in Generalized Extreme Value Regression: Modeling Annual Maximum Temperature
Next Article in Special Issue
Large-Time Behavior of Momentum Density Support of a Family of Weakly Dissipative Peakon Equations with Higher-Order Nonlinearity
Previous Article in Journal
Entropy-Based Tests for Complex Dependence in Economic and Financial Time Series with the R Package tseriesEntropy
Previous Article in Special Issue
The Buckling Operator: Inverse Boundary Value Problem
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Quasilinear Parabolic Equations Associated with Semilinear Parabolic Equations

1
Graduate School of Maritime Sciences, Kobe University, Kobe 658-0022, Japan
2
Department of Mathematics, École Normale Supérieure de Rennes, 35170 Bruz, France
3
Center for Mathematical Modeling and Data Science, Osaka University, Osaka 565-0871, Japan
*
Author to whom correspondence should be addressed.
Mathematics 2023, 11(3), 758; https://doi.org/10.3390/math11030758
Submission received: 31 December 2022 / Revised: 26 January 2023 / Accepted: 30 January 2023 / Published: 2 February 2023
(This article belongs to the Special Issue Modern Analysis and Partial Differential Equations, 2nd Edition)

Abstract

:
We formulate a quasilinear parabolic equation describing the behavior of the global-in-time solution to a semilinear parabolic equation. We study this equation in accordance with the blow-up and quenching patterns of the solution to the original semilinear parabolic equation. This quasilinear equation is new in the theory of partial differential equations and presents several difficulties for mathematical analysis. Two approaches are examined: functional analysis and a viscosity solution.

1. Introduction

The blow-up of the solution to the semilinear parabolic equation has been studied in detail. Here, we take
u t Δ u = u p in Ω × ( 0 , T ) , u ν Ω = 0 , u t = 0 = u 0 ( x )
for p > 1 , where Ω R N is a bounded domain with the smooth boundary Ω , ν denotes the outer unit normal vector on Ω , and 0 < u 0 = u 0 ( x ) C ( Ω ¯ ) . There exists a unique classical solution u = u ( x , t ) > 0 with the maximal existence time T = T max ( 0 , + ] . If T m a x < + , the solution blows up in finite time. Hence, it holds that
lim t T u ( · , t ) = + .
Here, we recall a few references related to the problem examined in this paper: the blow-up profile and the post-blow-up continuation of the solution. First, Masuda [1] studied the equation with the general nonlinear term
u t Δ u = f ( u ) in Ω × ( 0 , T ) , u ν Ω = 0 , u t = 0 = u 0 ( x ) .
By accepting the complex-valued t and u, the solution continues after the blow-up time is formulated. For the cases of f ( u ) = u 2 , u 2 + u , and e u , it is shown that this solution becomes two-fold in t > T unless u 0 is a constant.
Second, Baras–Cohen [2] and Lacey–Tzanetis [3] studied the solution u = u ( x , t ) 0 to
u t Δ u = f ( u ) in Ω × ( 0 , T ) , u Ω = 0 , u t = 0 = u 0 ( x )
for the non-negative nonlinearity f = f ( u ) 0 . Using the least integral solution, they formulate the notion of a complete blow-up of the solution, which means, roughly, that the solution u ( · , t ) becomes + everywhere in Ω for t > T . They showed this property under the appropriate conditions on f, Ω , and u 0 . Galaktionov–Vazquez [4] later refined the result such that if N 3 and
f ( u ) = u p , p ( ( N + 2 ) / ( N 2 ) , 1 + 6 / ( N 10 ) + ) ,
the radially symmetric unbounded L 1 -solution constructed by Ni–Sacks–Tzantzis [5] blows-up in finite time with the complete blow-up profile.
Third, Sakaguchi–Suzuki [6], motivated by their previous work [7] in the one space dimension using the Backlund transformation x = x ( u , t ) for u = u ( x , t ) , studied
u = u ( x , t ) C ( Ω ¯ × [ 0 , T ] ; ( , + ] ) ,
satisfying
D ( t ) = { x Ω u ( x , t ) = + } ¯ Ω , 0 t T
and the differential inequality
u t Δ u 0 in 0 t T ( Ω \ D ( t ) ) × { t }
in the sense of distributions. They obtained that then the N-dimensional Lebesgue measure of D ( t ) becomes zero for a.e. t. This result was later refined by Suzuki–Takahashi [8] as
0 T Cap 2 ( D ( t ) ) d t L N ( Ω ) 2 ,
where Cap 2 and L N denote the two-capacity and N-dimensional Lebesque measure, respectively. In particular, the Hausdorff dimension of D ( t ) is not greater than N 2 for a.e.t, if u = u ( x , t ) ( , + ] satisfies (4) and (6).
Here, we study the profile at t = T of the blow-up solution to (1). Our approach is based on the observation of Masuda [1], mentioned above, that the ordinary differential equation
d u d t = u p , u t = 0 = u 0 > 0 ,
uses the representation formula of the solution,
u ( t ) = 1 p 1 1 p 1 ( T t ) | T t | 1 p 1 1
for
T = u 0 ( p 1 ) p 1 .
Since (8) implies that
lim t T u ( t ) = + , lim t T u ( t ) = ,
we reach the idea of using 0 v < + , defined by
u 2 = v 1 p 1 ,
which satisfies
lim t T v ( t ) = 0 .
Since it holds that
v = 1 p 1 2 ( T t ) 2 , v t = 0 = v 0 1 p 1 2 T 2 > 0 ,
we obtain
d v d t = 2 ( p 1 ) w , t 0
for
w = v , 0 t T v , t > T
and
T = sup { t > 0 v ( s ) > 0 , 0 < s < t } .
The solution to (13) with (14) is not uniquely global in time. It is, however, unique to 0 t T , as we confirm below. Hence, we can use
d v d t = 2 ( p 1 ) v , v 0 , t 0
with
v t = 0 = v 0 > 0 ,
instead of (13) with (15), to detect the blow-up profile of the solution u = u ( x , t ) to (1) at t = T .
In fact, since the mapping
v [ 0 , + ) 2 ( p 1 ) v R
is non-increasing, the solution v = v ( t ) 0 to (16) and (17) is unique, although it is not Lispchitz continuous at v = 0 . More precisely, if v i , i = 1 , 2 , are the solutions to (16) it follows that
1 2 d d t v 1 v 2 2 2 = ( d v 1 d t d v 2 d t ) · ( v 1 v 2 ) 0
from this monotonicity.
Now we are able to recover u = u ( t ) for 0 t T using the solution v = v ( t ) from (16) and (17) for T > 0 defined by (15). Actually, it is given explicitly by
v ( t ) = 1 p 1 2 ( T t ) 2 , 0 t T 0 , t > T
for T > 0 , as defined by (9).
As for (1), it follows that
v t Δ v + 2 ( 2 p 1 ) p 1 | v | 2 + 2 ( p 1 ) v = 0 in Ω × ( 0 , T )
for v = v ( x , t ) > 0 , as defined by (11) and T = T max . We thus construct
0 v = v ( x , t ) C ( Ω ¯ × [ 0 , ) ) ,
satisfying
v t Δ v + 2 ( 2 p 1 ) p 1 | v | 2 + 2 ( p 1 ) v = 0 in Ω × ( 0 , )
with
v ν Ω = 0 , v t = 0 = v 0 ( x ) > 0
for v 0 = u 0 2 ( p 1 ) . Once this v is obtained, the value
T = sup { t > 0 min Ω ¯ v ( · , s ) > 0 , 0 < s < t }
coincides with the maximal existence time T = T max ( 0 , + ] of u = u ( · , t ) , and the function
u * = v ( · , T ) 1 2 ( p 1 ) C ( Ω ¯ ; ( 0 , + ] )
can stand for its blow-up profile at t = T if T = T max < + .
Multiplying v by a positive constant, we obtain the normal form of (18) and (19),
v t Δ v + γ | v | 2 + v = 0 , v 0 in Ω × ( 0 , T )
with
v ν Ω = 0 , v t = 0 = v 0 ( x ) > 0
for
γ = 2 ( 2 p 1 ) p 1 ,
which is the quasilinear parabolic equation studied in this paper. This equation is expected to clarify the blow-up patterns of the solution to (1). Table 1, below, summarizes several of the approaches to the blow-up problem of (1) that have been tried so far.
Surprisingly, the problems (20) and (21) are new in the theory of partial differential equations, and several technical difficulties arise in their mathematical analysis. Here, we take two approaches: the theory of functional analysis and that of a viscosity solution.
In the first approach, we use the approximate equation
v ε t Δ v ε + γ | v ε + ε | 2 + v ε = 0 , v ε 0 in Ω × ( 0 , )
with
v ε ν Ω = 0 , v ε t = 0 = v 0 ( x ) ,
assuming 0 < v 0 = v 0 ( x ) C θ ( Ω ¯ ) for 0 < θ < 1 , and try to pass the limit as ε 0 (Theorems 1 and 2).
In the second approach, we use w = v in (20), to obtain
( w 2 ) t Δ ( w 2 ) + γ | w | 2 + w = 0 , w 0 in Ω × ( 0 , T ) .
Equation (25) takes a different form from the equation studied in the standard theory of viscosity solutions, as referenced in Crandall–Ishii–Lions [10] and Koike [11], in which the comparison theorem, for example, guarantees the unique existence of the solution for the Dirichlet boundary condition
w Ω = 0 .
We show, however, that a part of this comparison principle is valid in (3) for γ 2 , which ensures a profile of the quenching of the solution to (3) for f ( u ) = u p with 0 < p < 1 (Theorem 5).
This paper is organized as follows: In Section 2 we derive a criterion for the convergence of the approximate solution defined by (23) and (24). Section 3 is concerned with the elliptic part of (25) and (26),
Δ ( w 2 ) + γ | w | 2 + w = 0 , w 0 in Ω
with
w Ω = 0 ,
and then, Section 4 deals with (25) and (26). Section 5 is devoted to the discussion, together with several examples of the nonlinearity f ( u ) in (2) or (3), to which our theorems are applicable.
  • List of Symbols
1.
C m ( Ω ¯ ) , m = 0 , 1 , 2 , ; set of m-times the continuously differentiable functions, where Ω R n is an open set. C ( Ω ¯ ) = C 0 ( Ω ¯ ) .
2.
C ( Ω ¯ × [ 0 , T ] ; ( , + ] ) , T > 0 ; set of continuous functions on Ω ¯ × [ 0 , T ] with the value in ( , + ] .
3.
W m , p ( Ω ) , 1 p ; set of measurable functions in Ω , with its distributional derivatives in L p ( Ω ) up to m-th order. H m ( Ω ) = W m , 2 ( Ω ) .
4.
L l o c p ( Ω ¯ × [ 0 , T ) ) ; set of measureable functions belonging to L p ( K ) for any compact set K Ω ¯ × [ 0 , T ) .
5.
C 2 + θ , 1 + θ / 2 ( Ω ¯ × [ 0 , ) ) , 0 < θ < 1 ; set of continuously differentiable functions up to the second and the first orders with respect to x Ω ¯ and t [ 0 , ) , with their derivatives Holder remaining continuous with the exponents θ and θ / 2 , respectively.
6.
L p ( 0 , T ; W m , q ( Ω ) ) ; set of L p functions in t ( 0 , T ) with the values in W m , q ( Ω ) .
7.
M ( Q ¯ ) = C ( Q ¯ ) ; set of measures on Q ¯ = Ω ¯ × [ 0 , T ] .

2. Convergence of the Approximate Solution

This section is concerned with the convergence of the approximate solution { v ε } defined by (23) and (24) for γ > 0 . For the moment, we use T > 0 arbitrarily.
To construct this solution, we use a further approximation defined for 0 < δ 1 , that is,
v ε t δ Δ v ε δ + γ | v ε δ + ε | 2 + ( v ε δ ) + + δ 2 = 0 , v ε δ > ε / 2 in Ω × ( 0 , T )
with
v ε δ ν Ω = 0 , v ε δ t = 0 = v 0 ( x ) ,
There is a local-in-time classical solution to (29) and (30) denoted by
ε / 2 < v ε δ = v ε δ ( x , t ) C 2 + θ , 1 + θ / 2 ( Ω ¯ × [ 0 , T ) ) ,
for 0 < T 1 , where 0 < θ < 1 .
The spatially homogeneous part of (29) and (30), on the other hand, is described by
d v d t = v + + δ 2 , v t = 0 = v 0 > 0 .
Its solution is given explicitly by
v ( t ) = δ 2 + 1 4 ( t δ t ) 2 , 0 t t δ 2 δ δ t + δ ( t δ 2 δ ) , t > t δ 2 δ ,
where
t δ = 2 v 0 + δ 2 .
Hence, we obtain
ε / 2 < v δ ( t ) v ε δ ( x , t ) v + δ ( t ) , ( x , t ) Ω ¯ × [ 0 , T ε δ )
by the comparison theorem, where
v ± δ ( t ) = δ 2 + 1 4 ( t ± δ t ) 2 , 0 t t ± δ 2 δ δ t + δ ( t ± δ 2 δ ) , t > t ± δ 2 δ , T ε δ = t δ 2 δ + δ 1 ε / 2 ,
and
t ± δ = 2 v 0 ± + δ 2 , v 0 + = max Ω ¯ v 0 , v 0 = min Ω ¯ v 0 > 0 ,
together with the existence of v ε δ = v ε δ ( x , t ) on Ω ¯ × [ 0 , T ε δ ) .
To make δ 0 with ε > 0 fixed, we use an estimate uniform in δ , that is,
v ε δ C 2 + θ , 1 + θ / 2 ( Ω ¯ × [ 0 , T ε δ ) ) C ε .
This estimate follows from the standard theory of quasilinear parabolic equations (Chapters IV and IV of [12]) inside Ω and a reflextion argument on Ω , as in Chapter 2 of [9] and Appendix A of [13]. We thus obtain v ε = v ε ( x , t ) C 2 + θ , 1 + θ / 2 ( Ω ¯ × [ 0 , ) ) satisfying
v ε t Δ v ε + γ | v ε + ε | 2 + v ε + = 0 , v ε ε / 2 in Ω × ( 0 , )
with
v ε ν Ω = 0 , v ε t = 0 = v 0 ( x ) .
Here, the maximum principle ensures v ε 0 , and there arises a solution
0 v ε = v ε ( x , t ) C 2 + θ , 1 + θ / 2 ( Ω ¯ × [ 0 , ) )
to (23) and (24), satisfying
v ε ( · , t ) v 0 .
Now we show a criterion for this { v ε } to converge to a solution v = v ( x , t ) to (20) as ε 0 , such that
0 v L l o c ( Ω ¯ × [ 0 , ) ) , v L 2 ( 0 , T ; H 1 ( Ω ) )
for the case of γ 2 .
First, Inequality (32) implies
v ε ( · , t ) v 0 ,
while
d d t Ω v ε d x + γ v ε + ε 2 2 0
implies
0 T v ε + ε 2 2 d t γ 1 v 0 1 .
Second, the approximate solution v ε ( x , t ) is monotone in ε for each ( x , t ) Ω ¯ × [ 0 , ) . To see this property, we make a further approximation, using
φ δ σ ( v ) φ δ ( v ) locally and uniformly in v R , σ 0 ,
where φ δ σ C 1 ( R ) and φ δ ( v ) = v + + δ 2 , that is, v ε σ , δ = v ε σ , δ ( x , t ) satisfying
v ε t σ , δ Δ v ε σ , δ + γ v ε σ , δ + ε 2 + φ ε σ , δ ( v ε σ , δ ) = 0 , v ε σ , δ > ε / 2 in Ω × ( 0 , T ε σ , δ ) ,
with
v ε σ , δ ν Ω = 0 , v ε σ , δ t = 0 = v 0 ( x ) .
It follows that T ε σ , δ T ε δ , and
v ε σ , δ v ε δ locally and uniformly on Ω ¯ × [ 0 , T ε δ ) , σ 0 .
There is
z ε σ , δ = v ε σ , δ ε ,
and it holds that
z ε t σ , δ Δ z ε σ δ + 2 γ ( v ε σ , δ + ε ) 1 / 2 · 1 2 ( v ε σ , δ + ε ) 1 / 2 ( z ε δ + 1 ) + [ ( φ δ σ ) ( v ε σ , δ ) ] z ε σ , δ = 0 in Ω × ( 0 , T ε σ , δ )
with
z ε σ , δ ν Ω = 0 , z ε σ , δ t = 0 = 0 .
Then, we obtain
z ε t σ , δ Δ z ε σ , δ + γ ( v ε σ , δ + ε ) 1 / 2 · [ ( v ε σ , δ + ε ) 1 / 2 z ε σ , δ ] + [ ( φ δ σ ) ( v ε σ , δ ) ] z ε σ , δ 0 in Ω × ( 0 , T ε γ , δ )
by
( v ε σ , δ + ε ) 1 / 2 · ( v ε σ , δ + ε ) 1 / 2 = ( v ε σ , δ + ε ) | ( v ε σ , δ + ε ) 1 / 2 | 2 0 ,
which implies
z ε σ , δ = v ε σ , δ ε 0 o n Ω ¯ × [ 0 , T ε σ , δ ) .
Then, it arises that
v ε σ , δ ( x , t ) v ε σ , δ ( x , t ) , ( x , t ) Ω ¯ × [ 0 , T ε σ , δ ) , ε > ε > 0
and, hence,
v ε ( x , t ) v ε ( x , t ) , ( x , t ) Ω ¯ × [ 0 , ) , ε > ε > 0
by sending σ 0 , and then δ 0 . Thus it arises that the pointwise monotone convergence
lim ε 0 v ε ( x , t ) = v ( x , t ) 0 , ( x , t ) Ω ¯ × [ 0 , T ) .
Letting
w ε = v ε + ε , Q T = Ω × ( 0 , T ) ,
we obtain
w ε L ( Q T ) + w ε L 2 ( Q T ) C
via (36) and (37), and the monotone convergence
lim ε 0 w ε ( x , t ) = w ( x , t ) v ( x , t ) , ( x , t ) Ω ¯ × [ 0 , T )
by (38). Hence, it follows that
0 v L ( Q T ) , v L 2 ( 0 , T ; H 1 ( Ω ) )
and
w lim ε 0 w ε = w in L 2 ( 0 , T ; H 1 ( Ω ) ) ,
where w lim denotes the weak convergence. Then, we have the following theorem.
Theorem 1. 
If γ 2 and w C ( Q T ¯ ) , the limit v C ( Q T ¯ ) in (38) is a solution to (20) and (21) in the sense of distributions.
We begin with the following lemma.
Lemma 1. 
If the convergence (41) is strong,
s lim ε 0 w ε = w i n L 2 ( 0 , T ; H 1 ( Ω ) ) ,
the above v = v ( x , t ) in (40) is a solution to (20) and (21) in the sense of distributions.
Proof. 
Assumption (42) implies
s lim ε 0 v ε = v in L 1 ( 0 , T ; W 1 , 1 ( Ω ) )
by
v ε = 2 w ε w ε .
It also implies
s lim ε 0 v ε + ε = v in L 2 ( 0 , T ; H 1 ( Ω ) )
by
v ε + ε = ( w ε 2 + ε ) 1 / 2 w ε w ε , 0 ( w ε 2 + ε ) 1 / 2 w ε 1 ,
and the monotone convergence (39), as in
v ε + ε v = ( w ε 2 + ε ) 1 / 2 w ε w ε w = { ( w ε 2 + ε ) 1 / 2 w ε 1 } w ε + ( w ε w ) .
Given
φ C 2 ( Ω ¯ ) , φ ν Ω = 0 ,
we obtain
d d t Ω v ε φ d x + ( v ε , φ ) + Ω γ | v ε + ε | 2 + v ε φ d x = 0
by (31), (23), and (24), where ( , ) denotes the L 2 inner product. Then, (42)–(44) imply
d d t Ω v φ d x + ( v , ϕ ) + Ω ( γ | v | 2 + v ) φ d x = 0
in the sense of distributions in t. In particular, the mapping
t [ 0 , T ) Ω v φ d x
is absolutely continuous, and it holds that
Ω v φ d x t = 0 = Ω v 0 φ d x .
Then, the result follows. □
Equation (31) with (23) implies
w ε t Δ w ε = g ε in Q T
with
w ε ν Ω = 0 , w ε t = 0 = w 0 ( x ) > 0
for w 0 = v 0 1 / 2 and
g ε = 2 ( γ 2 ) | w ε | 2 + v ε 2 v ε + ε 0
if γ 2 . Then, we obtain
d d t Ω w ε d x = Ω g ε d x 0 ,
and, therefore,
g ε L 1 ( Q T ) v 0 1 / 2 1
by w ε 0 . Hence, there is a subsequence, denoted by the same symbol, such that
g ε μ M ( Q T ¯ ) = C ( Q T ¯ )
in the sense of measures.
Remark 1. 
Inequality (48) implies
w C ( [ 0 , + ) , L 1 ( Ω ) )
according to (41) and the L 1 -compactness property of the heat equation [14]. This inequality also ensures
s lim ε 0 w ε = w i n L p ( Q T ) , 1 p < N + 2 N s lim ε 0 w ε = w i n L q ( Q T ) , 1 q < N + 2 N + 1
by [15].
Lemma 2. 
If γ 2 and w C ( Q T ¯ ) , it holds that
0 T ( w ε w ) 2 2 d t w ε w , μ + 1 2 v 0 + ε v 0 2 2 .
Proof. 
We have
( w ε w ε ) t Δ ( w ε w ε ) = g ε + g ε in Q T
with
ν ( w ε w ε ) Ω = 0 , ( w ε w ε ) t = 0 = 0 .
Then, it follows that
1 2 d d t w ε w ε 2 2 + ( w ε w ε ) 2 2 = ( w ε w ε , g ε + g ε ) ( w ε w ε , g ε )
for 0 < ε < ε , and, therefore,
0 T ( w ε w ε ) 2 2 d t 0 T ( w ε w ε , g ε ) d t + 1 2 v 0 + ε v 0 + ε 2 2 .
If w C ( Q T ¯ ) , the monotone convergence (39) implies
lim ε 0 w ε = w uniformly on Q T ¯ .
by Dini’s theorem. Then, it holds that
lim ε 0 0 T ( w ε w ε , g ε ) d t = w ε w , μ ,
by (41). Hence, (49) follows from
lim inf ε 0 0 T ( w ε w ε ) 2 2 d t 0 T ( w ε w ) 2 2 d t .
We are now able to give the following proof.
Proof of Theorem 1. 
The result follows from (49) and (51). □
A variant of Theorem 1 is the following theorem.
Theorem 2. 
Let 2 γ < 4 and assume the existence of w ^ ε , w * , w * C ( Q T ¯ ) such that
w ε w ^ ε , w ^ ε w * u n i f o r m l y   o n   Q T ¯ ,
and
w w * o n   Q T ¯ , w * w * , μ = 0 .
Then, it holds that (42), and hence v in (38), satisfies (12), (20), and (21) in the sense of distributions.
We use the following lemma to prove this theorem.
Lemma 3. 
If γ 2 , it holds that
0 T ( w ε w ) 2 2 d t w ε , μ Q T γ 2 2 | w | 2 + w 2 d x d t + 1 2 v 0 + ε v 0 2 2 .
Proof. 
In (50), we have
( w ε , g ε ) = Ω γ 2 2 | w ε | 2 + 1 2 v ε v ε + ε w ε d x Ω γ 2 2 | w ε | 2 + w ε 2 d x
by (47), and, therefore, (41) ensures
lim inf ε 0 0 T ( w ε , g ε ) d t Q T γ 2 2 | w | 2 + w 2 d x d t .
Then, we obtain (55) by (52). □
We conclude this theorem with the following proof.
Proof of Theorem 2. 
In (55), we have
w ε , μ = w ε w ^ ε , μ + w ^ ε , μ w ^ ε , μ = w * , μ + o ( 1 ) = w * , μ + o ( 1 ) .
It holds, furthermore, that
w * , μ = w * , g ε + o ( 1 ) w , g ε + o ( 1 )
and
w , g ε = Q T γ 2 2 | w ε | 2 w ε w + 1 2 v ε v ε + ε w d x d t Q T γ 2 2 | w ε | 2 + w 2 d x d t + o ( 1 )
by (47). Then, it follows that
0 T ( w ε w ) 2 2 d t γ 2 2 Q T | w ε | 2 | w | 2 d x d t + o ( 1 ) = γ 2 2 0 T ( w ε w ) 2 2 d t + o ( 1 )
from (41), which implies (42) by γ < 4 . □

3. Comparison Theorem for the Elliptic Equation

In this section, we study the viscosity solutions to (27) with (28). To introduce this notion, let Ω R N be a bounded domain, and let U S C ( Ω ) (resp., L S C ( Ω ) ) be the set of upper semi-continuous (resp., lower semi-continuous) functions in Ω . Given w : Ω R , we define w * and w * by
w * ( x ) = lim sup y x w ( y ) , w * ( x ) = lim inf y x w ( y )
for x Ω . If w C ( Ω ) , it holds that w = w * = w * .
Remark 2. 
We have w * U S C ( Ω ) if and only if w * ( x ) < for any x Ω . Similarly, we have w * L S C ( Ω ) if and only if w * ( x ) > for any x Ω .
Remark 3. 
If w : Ω ¯ R , the above w * and w * are extended as
w * ( x ) = lim sup y Ω ¯ x w ( x ) , w * ( x ) = lim inf y Ω ¯ x w ( y )
for x Ω ¯ . The same properties as in Remark 2 are valid for w * and w * .
Using
F ( r , p , X ) = t r X + γ | p | 2 + r , r R , p R N , X S N ( R ) ,
we write (27) as
F ( w , w , 2 w 2 ) = 0 in Ω ,
where
S N = { A M N ( R ) A T = A } ,
M N ( R ) is the set of N × N matrix with real entries, and A T denotes the transpose matrix of A.
Since Equation (57) takes a different form when treated with the standard theory of visicosity solution,
F ( w , w , 2 w ) = 0 in Ω ,
we begin with the definition of its viscosity solution.
Definition 1. 
Let w = w ( x ) be a function defined in Ω.
(1) 
We say that if w in w * U S C ( Ω ) is a viscosity subsolution to (27) if w * φ attains a local maximum 0 at x Ω for φ C 2 ( Ω ) , then it holds that
F ( φ , φ , 2 φ 2 ) 0 a t   x .
(2) 
We say that if w in w * L S C ( Ω ) is a viscosity supersolution to (27) if w * φ attains a local minimum 0 at x 0 Ω for φ C 2 ( Ω ) , then it holds that
F ( φ , φ , 2 φ 2 ) 0 a t   x 0 .
(3) 
We say that if w is a viscosity solution to (27) if it is a viscosity subsolution and a viscosity supersolution to (27).
Remark 4. 
It is obvious that if w C 2 ( Ω ) is a viscosity subsolution to (27), then it holds that
F ( w , w , 2 w 2 ) 0 i n   Ω ,
and hence, it is a classical subsolution. Similarly, if w C 2 ( Ω ) is a viscosity supersolution (resp., solution) to (27), it is a classical supersolution (resp., solution).
Even if w C 2 ( Ω ) is a classical solution to (27),
F ( w , w , 2 w 2 ) = 0 in Ω .
On the other hand, it is not necessarily a viscosity subsolution, nor a viscosity supersolution, although F in (56) is elliptic:
X , Y S N , X Y F ( r , p , Y ) F ( r , p , X ) , ( r , p ) R × R N .
Despite this, we have the following fact.
Proposition 1. 
If 0 w = w ( x ) C 2 ( Ω ) is a classical subsolution to (27), it is a viscosity subsolution. If 0 < w = w ( x ) C 2 ( Ω ) is a classical supersolution to (27), it is a viscosity supersolution.
To prove the above fact, we recall, for the moment, a fundamental fact used in the theory of viscosity solutions.
Given w : Ω R and x Ω , we thus put
J 2 , + w ( x ) = { ( p , X ) R N × S N w * ( x + h ) w * ( x ) p , h + 1 2 X h , h + o ( | h | 2 ) a s h 0 } .
Given 0 < r 1 and ( p , X ) R N × S N , furthermore, we put
Φ r + ( x , p , X , w ) = { φ C 2 ( B r ( x ) ) w * φ a t t a i n s   t h e   m a x i m u m   0   a t   x , φ ( x ) = p , 2 φ ( x ) = X } .
Let
Φ + ( x , p , X ; w ) = 0 < r 1 Φ r + ( x , p , X ; w ) .
The following fact is proven in Koike’s work ([11], Proposition 2.6).
Lemma 4. 
Any x Ω and ( p , X ) J 2 , + w ( x ) admit φ Φ + ( x , p , X ; w ) .
Remark 5. 
We define Φ ( x , p , X ; w ) similarly, using
J 2 , w ( x ) = { ( p , X ) R N × S N w * ( x + h ) w * ( x ) p , h + 1 2 X h , h + o ( h 2 ) , h 0 } .
Then, an analogous result to the above lemma holds.
An immediate consequence is the following fact, analogous to the work of Koike ([11], Corollary 2.3), in which
j 2 , + w ( x ) = ( φ ( x ) , 2 φ ( x ) 2 ) ( p , X ) J 2 , + w ( x ) , φ Φ + ( x , p , X ; w ) .
Corollary 1. 
The function w is a viscosity subsolution to (27) if and only if
F ( w * ( x ) , q , Y ) 0 , x Ω , ( q , Y ) j 2 , + w ( x ) .
We are ready to give the following proof. Note that (60) does not imply (61) without w ( y ) 0 , and, similarly, (62) does not imply (63) without φ ( y ) 0 .
Proof of Proposition 1. 
Let 0 w C 2 ( Ω ) be a classical subsolution:
F ( w , w , 2 w ) 0 i n   Ω .
Observe that w = w * = w * in Ω .
Fix x Ω and use ( q , Y ) j 2 , + w ( x ) arbitrarily. Then, there are ( p , X ) J 2 , + w ( x ) and φ Φ ( x , p , X ; w ) such that
q = φ ( x ) , Y = 2 φ ( x ) 2 .
Next, we obtain
w ( y ) φ ( y ) w ( x ) φ ( x ) = 0 , | y x | 1 ,
and hence
w ( y ) 2 φ ( y ) 2 , | y x | 1 , w ( x ) 2 = φ ( x ) 2
by w ( y ) 0 . It follows, therefore,
( w ( x ) φ ( x ) ) = 0 , 2 ( w ( x ) 2 φ ( x ) 2 ) 0 ,
which implies
F ( w ( x ) , q , Y ) = F ( φ ( x ) , φ ( x ) , 2 φ ( x ) 2 ) F ( w ( x ) , w ( x ) , 2 w ( x ) 2 ) 0
by (59). Hence, 0 w = w ( x ) C 2 ( Ω ) is a viscosity subsolution.
The proof of the latter part is similar. In fact, if 0 < w C 2 ( Ω ) and φ C 2 ( Ω ) satisfy
w ( y ) φ ( y ) w ( x ) φ ( x ) = 0 , | y x | 1
then it arises that
w ( y ) 2 φ ( y ) 2 , | y x | 1 , w ( x ) 2 = φ ( x ) 2
by w ( x ) > 0 . Hence, if 0 < w = w ( x ) C 2 ( Ω ) is a classical supersolution, then it is a viscosity supersolution. □
Here, we show the following fact.
Theorem 3. 
Assume γ 2 . Let w 0 (resp., z 0 ) be a viscosity subsolution (resp., supersolution) to (27) defined on Ω ¯ , and put
Ω 0 = { x Ω | z * ( x ) > 0 } .
Then, if
sup Ω 0 ( w * z * ) 0 , sup Ω ¯ 0 z < +
it holds that
sup Ω ¯ 0 ( w * z * ) 0 .
Remark 6. 
The set Ω 0 is open in Ω because z * is lower semicontinuous on Ω ¯ according to the same reason z * = 0 in Ω 0 \ Ω .
For the proof of this theorem, we use the following notations used in the theory of viscosity solutions; namely, given w : Ω R , let
J ¯ 2 , + w ( x ) = { ( p , X ) R N × S N ( x k , p k , X k ) Ω × R N × S N , k = 1 , 2 , , lim k ( x k , w * ( x k ) , p , X ) = ( x , w * ( x ) , p , X ) , ( p k , X k ) J 2 , + w ( x k ) } ,
and
J ¯ 2 , w ( x ) = { ( p , X ) R N × S N ( x k , p k , X k ) Ω × R N × S N , k = 1 , 2 , , lim k ( x k , w * ( x k ) , p , X ) = ( x , w * ( x ) , p , X ) , ( p k , X k ) J 2 , w ( x k ) }
Proof of Theorem 3. 
Assume the contrary,
sup Ω ¯ 0 ( w * z * ) = θ > 0 .
We take 0 < ρ < 1 , satisfying
sup Ω 0 ( w * ρ z * ) ( 1 ρ ) sup Ω 0 z * ( 1 ρ ) sup Ω ¯ 0 z * θ 3 ,
and put
θ ˜ = sup Ω ¯ 0 ( w * ρ z * ) θ .
Let ε > 0 and ( x ^ ε , y ^ ε ) Ω ¯ 0 × Ω ¯ 0 be the maximum point of
Φ ε ( x , y ) = w * ( x ) ρ z * ( y ) 1 2 ε | x y | 2 , ( x , y ) Ω ¯ 0 × Ω ¯ 0 .
Since
Φ ε ( x ^ ε , x ^ ε ) Φ ε ( x ^ ε , y ^ e ) , z * ( y ^ ε ) 0
it holds that
1 2 ε | x ^ ε y ^ ε | 2 ρ z * ( x ^ ε ) ρ sup Ω ¯ 0 z * ,
and hence,
lim ε 0 | x ^ ε y ^ ε | = 0 .
A subsequence, therefore, admits x 0 Ω ¯ 0 such that
lim ε 0 x ^ ε = lim ε 0 y ^ ε = x 0 .
Since Φ ε ( x 0 , x 0 ) Φ ε ( x ^ ε , y ^ ε ) , it holds that
ρ ( z * ( y ^ ε ) z * ( x 0 ) ) w ( x ^ ε ) w * ( x 0 ) ,
and therefore,
0 lim inf ε 0 ρ ( z * ( y ^ ε ) z * ( x 0 ) ) lim inf ε 0 ( w * ( x ^ ε ) w * ( x 0 ) ) lim sup ε 0 ( w * ( x ^ ε ) w * ( x 0 ) ) 0
by the semicontinuity. We thus obtain
lim ε 0 w * ( x ^ ε ) = w * ( x 0 ) , lim ε 0 z * ( y ^ ε ) = z * ( x 0 ) ,
which implies
lim ε 0 1 ε | x ^ ε y ^ ε | 2 = 0
by Φ ε ( x 0 , x 0 ) Φ ε ( x ^ ε , y ^ ε ) .
Since
sup Ω 0 ( w * ρ z * ) θ / 3 < θ ˜ Φ ε ( x ^ ε , y ^ ε ) ,
Equalities (64) and (65) imply x 0 Ω 0 , which means z * ( x 0 ) > 0 . Noting
0 < θ ˜ Φ ε ( x ^ ε , y ^ ε ) w * ( x ^ ε ) ρ z * ( y ^ ε ) ,
thus we obtain
w * ( x ^ ε ) > ρ z * ( y ^ ε ) > 0 , 0 < ε 1 .
Then, H. Ishii’s lemma, as referenced in Koike ([11], Appendix A), guarantees X ^ ε , Y ^ ε S N satisfying
( p ^ ε , X ^ ε ) J ¯ 2 , + w ( x ^ ε ) , ( p ^ ε , Y ^ ε ) J ¯ 2 , ρ z ( y ^ ε )
and
3 ε I O O I X ^ ε O O Y ^ ε 3 ε I I I I
for p ^ ε = ( x ^ ε y ^ ε ) / ε . Note that the second relation of (67) means
( p ^ ε / ρ , Y ^ ε / ρ ) J ¯ 2 , z ( y ^ ε ) .
Then, Lemma 4 and Remark 5 ensure
( x k , p k , X k , ϕ k ) , ( y k q k , Y k , ψ k ) Ω × R N × S N , k = 1 , 2 , ,
satisfying
ϕ k Φ + ( x k , p k , X k ; w ) , ψ k Φ ( y k , q k , Y k , z )
and
lim k ( x k , w * ( x k ) , p k , X k ) = ( x ^ ε , w * ( x ^ ε ) , p ^ ε , X ^ ε ) ,
lim k ( y k , z * ( y k ) , q k , Y k ) = ( y ^ ε , z * ( y ^ ε ) , p ^ ε / ρ , Y ^ ε / ρ ) .
Since w (resp., z) is a viscosity subsolution (resp., supersolution) to (27), it holds that
F ( ϕ k , ϕ k , 2 ϕ k 2 ) 0 a t x k , F ( ψ k , ψ k , 2 ψ k 2 ) 0 at   y k .
Here, we use
t r 2 ϕ 2 = Δ ϕ 2 = 2 | ϕ | 2 + 2 ϕ Δ ϕ
valid for ϕ C 2 ( Ω ) , to obtain
2 ϕ k tr X k + ( γ 2 ) | p k | 2 + ϕ k 0 at   x k , 2 ψ k tr ( Y k ) + ( γ 2 ) | q k | 2 + ψ k 0 at   y k
by (69). Sending k , it arises that
2 w * ( x ^ ε ) tr X ^ ε + ( γ 2 ) | p ^ ε | 2 + w * ( x ^ ε ) 0 , 2 z * ( y ^ ε ) tr ( Y ^ ε ρ ) + γ 2 ρ 2 | p ^ ε | 2 + z * ( y ^ ε ) 0 ,
from (70) and (71), which implies
2 tr X ^ ε + γ 2 w * ( x ^ ε ) | p ^ ε | 2 + 1 0 , 2 tr ( Y ^ ε ) + γ 2 ρ z * ( y ^ ε ) | p ^ ε | 2 + ρ 0
by (66).
Since (68) implies
tr ( X ^ ε + Y ^ ε ) 0 ,
we obtain
1 w * ( x ^ ε ) 1 ρ z * ( y ^ ε ) ( γ 2 ) | p ^ ε | 2 2 tr ( X ^ ε + Y ^ ε ) + ρ 1 < 0 ,
which is in contradiction to γ 2 and (66). Thus, the result follows. □
A direct consequence of Theorem 3 is the following fact based on its uniqueness and regularity.
Corollary 2. 
Assume γ 2 . Let w 0 and z 0 be bounded viscosity solutions to (27), defined on Ω ¯ , and put
Ω 0 = { x Ω | w * ( x ) > 0 } { x Ω | z * ( x ) > 0 } .
Then, if w * = w * = z * = z * on Ω 0 , it follows that u = v C ( Ω ¯ 0 ) .
Proof. 
This result is similar to the standard case. In fact, the assumption implies
w * z * z * w * w * o n   Ω ¯ 0 .
The following theorem is obvious if w C ( Ω ¯ ) .
Theorem 4. 
Let w 0 be a bounded viscosity solution to (27), defined on Ω ¯ with γ 2 satisfying
w * = w * = 0 o n   ( Ω \ Ω 0 ) ,
and put
Ω 0 = { x Ω w * ( x ) > 0 } .
Then, it follows that
w * = w * = 0 o n   Ω \ Ω 0 ¯ .
Proof. 
Assume the contrary,
sup Ω \ Ω 0 ¯ w * > 0 ,
and let x 0 Ω \ Ω 0 ¯ be a maximum point of w * on Ω \ Ω 0 ¯ . Then, we obtain x 0 Ω \ Ω ¯ 0 by (72). Since Ω \ Ω ¯ 0 is open, we find r 0 > 0 such that w * ( x ) w * ( x 0 ) for all x B r 0 ( x 0 ) .
Let φ ( x ) = w * ( x 0 ) . It is obvious that w * φ attains a local maximum at x 0 . Since w 0 is a viscosity subsolution to (27), it arises that
0 F ( φ , φ , 2 φ 2 ) = w * ( x 0 ) > 0 at   x 0 ,
which is a contradiction. Thus, w * 0 on Ω \ Ω 0 ¯ . Since w 0 , it follows (74). □

4. Comparison Theorem to (25)

The viscosity solutions to (25) are treated similarly. Recall F ( r , p , X ) for ( r , p , X ) R × R n × S N in (56).
Definition 2. 
Let w : Q T = Ω × ( 0 , T ) R .
(1) 
We say that w with w * < + in Q T is a viscosity subsolution to (25), provided that if w * φ attains a local maximum 0 at ( x , t ) Q T for φ C 2 ( Q T ) , then it holds that
( φ 2 ) t + F ( φ , φ , 2 φ 2 ) 0 a t   ( x 0 , t 0 ) .
(2) 
We say that w in w * > in Q T is a viscosity supersolution to (25), provided that if w * φ attains a local minimum 0 at ( x , t ) Q T for φ C 2 ( Q T ) , then it holds that
( φ 2 ) t + F ( φ , φ , 2 φ 2 ) 0 a t   ( x , t ) .
(3) 
We say that w is a viscosity solution to (25) if it is a viscosity subsolution and a viscosity supersolution to (25).
Similarly to Remark 4 and Proposition 1, we obtain the following fact.
Proposition 2. 
If w C 2 ( Q T ) is a viscosity subsolution (resp., supersolution) to (25), it is a classical subsolution (resp., supersolution). If 0 w C 2 ( Q T ) is a classical subsolution to (25), conversely, it is a viscosity subsolution. If 0 < w C 2 ( Q T ) is classical supersolution to (25), finally, it is a viscosity supersolution.
Given ( x , t ) Q T and r > 0 , we use the following sets, where Q r ( x , t ) = B r ( x ) × ( t r , t + r ) :
P 2 , + w ( x , t ) = { ( p , τ , X ) R × R N × S N w * ( x + h , t + k ) w * ( x , t ) τ k + p , h + 1 2 X h , h + o ( | k | + | h | 2 ) a s ( h , k ) ( 0 , 0 ) } .
P 2 , w ( x , t ) = { ( p , τ , X ) R × R N × S N w * ( x + h , t + k ) w * ( x , t ) τ k + p , h + 1 2 X h , h + o ( | k | + | h | 2 ) a s ( k , h ) ( 0 , 0 ) } .
P ¯ 2 , + w ( x , t ) = { ( p , τ , X ) R × R N × S N ( x k , p k , τ k , X k ) Q T × R N × S N , k = 1 , 2 , , lim k ( x k , τ k , w * ( x k , τ k ) , p k , X k ) = ( x , τ , w * ( x , τ ) , p , X ) , ( p k , τ k , X k ) P 2 , + w ( x k , t k ) } .
P ¯ 2 , w ( x , t ) = { ( p , τ , X ) R × R N × S N ( x k , τ k , p k , X k ) Q T × R N × S N , k = 1 , 2 , , lim k ( x k , τ k , w * ( x k , t k ) , p k , X k ) = ( x , τ , w * ( x , τ ) , p , X ) , ( p k , τ k , X k ) P 2 , w ( x k , t k ) } .
Φ r + ( x , t , p , τ , X ; w ) = { φ C 2 ( Q r ( x , t ) ) ) w * φ a t t a i n s   i t s   m a x i m u m   0   a t   ( x , t ) , φ ( x , t ) = p , φ t ( x , t ) = τ , 2 φ ( x , t ) = X } .
Φ r ( x , t , p , τ , p , X ; w ) = { φ C 2 ( Q r ( x , t ) ) w * φ a t t a i n s   i t s   m i n i m u m   0   a t   ( x , t ) , φ ( x , t ) = p , φ t ( x , t ) = τ , 2 φ ( x , t ) = X } .
Φ ± ( x , t , p , τ , p , X ; w ) = 0 < r 1 Φ r ± ( x , t , p , τ , X ; w ) .
p 2 , ± w ( x , t ) = { ( φ ( x , t ) , φ t ( x , t ) , 2 φ ( x , t ) 2 ) φ Φ ± ( x , t , p , τ , p , X ; w ) , ( p , τ , X ) P 2 , + w ( x , t ) } .
Similarly to Lemma 4, we obtain the following fact.
Lemma 5. 
If w is locally bounded in Q T , each ( x , t ) Q T and ( p , τ , X ) P 2 , ± w ( x , t ) admits φ Φ ± ( x , t , p , τ , X ; w ) .
The comparison principle for the viscosity solution to (25) is described as follows. Although it is proven similarly to Theorem 3, we show the proof for completeness.
Theorem 5. 
Assume γ 2 . Let w 0 (resp., z 0 ) be a viscosity subsolution (resp., supersolution) to (25) defined on Q T ¯ , and put
Q 0 = { ( x , t ) Q T z * ( x , t ) > 0 } .
Then, if
sup Q 0 \ ( Ω ¯ × { T } ) ( w * z * ) 0 , sup Q T ¯ w < ,
it holds that
sup Q 0 ¯ \ ( Ω ¯ × { T } ) ( w * z * ) 0 .
Proof. 
Assuming the contrary,
sup Q 0 ¯ ( w * z * ) = θ > 0 ,
we take 0 < ρ , γ < 1 , satisfying
sup Q 0 \ ( Ω ¯ × { T } ) ( w * ρ z * 2 γ T t ) θ 3 .
It is obvious that
sup Q 0 ¯ \ ( Ω ¯ × { T } ) ( w * ρ z * 2 γ T t ) θ ˜ θ .
Let
Φ ε ( x , t , y , s ) = w * ( x , t ) ρ z * ( y , s ) 1 2 ε { | x y | 2 + ( t s ) 2 } γ T t γ T s
be defined for ( x , t , y , s ) Q 0 ˜ × Q 0 ˜ , where ε > 0 and
Q 0 ˜ = Q 0 ¯ \ ( Ω ¯ × { T } ) .
Let, furthermore, ( x ^ ε , t ^ ε , y ^ ε , s ^ ε ) Q 0 ˜ × Q 0 ˜ be a maximum point of Φ ε . Then, the inequality
Φ ε ( x ^ ε , t ^ ε , x ^ ε , t ^ ε ) + Φ ε ( y ^ ε , s ^ ε , y ^ ε , s ^ ε ) 2 Φ ε ( x ^ ε , t ^ ε , y ^ ε , s ^ ε )
implies
lim ε 0 { ( t ^ ε s ^ ε ) 2 + | x ^ ε y ^ ε | 2 } = 0 ,
and, furthermore, it arises that
lim ε 0 ( x ^ ε , t ^ ε ) = lim ε 0 ( y ^ ε , s ^ ε ) = ( x 0 , t 0 ) Q 0 ¯ lim ε 0 w * ( x ^ ε , t ^ ε ) = w * ( x 0 , t 0 ) , lim ε 0 z * ( y ^ ε , s ^ ε ) = z * ( x 0 , t 0 ) , lim ε 0 1 ε { ( t ^ ε s ^ ε ) 2 + | x ^ ε y ^ ε | 2 } = 0 .
as in Theorem 3. Using the fact that ( x 0 , t 0 ) is also a maximum point of
w * ( x , t ) ρ z * ( x , t ) 2 γ T t on Q 0 ˜ ,
we obtain ( x 0 , t 0 ) Q 0 and also
w * ( x ^ ε , t ^ ε ) > ρ z * ( y ^ ε , s ^ ε ) > 0 , 0 < ε 1 .
Then, H. Ishii’s lemma guarantees X ^ ε , Y ^ ε S N such that
( p ^ ε , τ ε + γ ( T t ^ ε ) 2 , X ^ ε ) P ¯ 2 , + w ( x ^ ε , t ^ ε ) ,
ρ 1 ( p ^ ε , τ ε γ ( T s ^ ε ) 2 , Y ^ ε ) P ¯ 2 , z ( y ^ ε , s ^ ε ) ,
and
3 ε I O O I X ^ ε O O Y ^ ε 3 ε I I I I
for
τ ε = ( t ^ ε s ^ ε ) / ε , p ^ ε = ( x ^ ε y ^ ε ) / ε .
By the definition of P ¯ 2 , ± and Lemma 5, we have
( x k , t k , p k , X k , ϕ k ) , ( y k , s k , q k , Y k , ψ k ) , k = 1 , 2 , ,
satisfying
ϕ k Φ + ( x k , t k , p k , τ k , X k ; w ) , ψ k Φ ( y k , s k , q k , κ k , Y k ; z )
and
lim k ( x k , t k , w * ( x k , t k ) , p k , τ k , X k ) = ( x ^ ε , t ^ ε , w * ( x ^ ε , t ^ ε ) , p ε , τ ε + γ ( T t ^ ε ) 2 , X ^ ε ) ,
lim k ( y k , s k , z * ( y k , s k ) , q k , κ k , Y k ) = ( y ^ ε , s ^ ε , z * ( y ^ ε , s ^ ε ) , p ε ρ , τ ^ ε ρ γ ρ ( T s ^ ε ) 2 , Y ^ ε ρ ) ,
which implies
( φ k 2 ) t + F ( ϕ k 2 , ϕ k , 2 ϕ k 2 ) 0 at ( x k , t k ) , ( ψ k 2 ) t + F ( ψ k , ψ k , 2 ( ψ k ) 2 ) 0 at ( y k , s k ) ,
because w (resp., z) is a viscosity subsolution (resp., supersolution) to (27). Then, there holds that
2 ϕ k τ k 2 ϕ k tr X k + ( γ 2 ) | p k | 2 + ϕ k 0 at ( x k , t k ) , 2 ψ k κ k 2 ψ k tr ( Y k ) + ( γ 2 ) | q k | 2 + ψ k 0 at ( s k , y k ) .
by (80).
Sending k , we obtain
2 w * ( x ^ ε , t ^ ε ) { τ ^ ε + γ ( T t ^ ε ) 2 } 2 w * ( x ^ ε , t ^ ε ) tr X ^ ε + ( γ 2 ) | p ^ ε | 2 + w * ( x ^ ε , t ^ ε ) 0 , 2 z * ( y ^ ε , s ^ ε ) { τ ^ ε ρ γ ρ ( T t ^ ε ) 2 } 2 z * ( y ^ ε , s ^ ε ) tr ( Y ^ ε ρ ) + γ 2 ρ 2 | p ^ ε | 2 + v ( y ^ ε , s ^ ε ) 0 .
by (81) and (82), which implies
2 { τ ^ ε + γ ( T t ^ ε ) 2 } 2 tr X ^ ε + γ 2 w * ( x ^ ε , t ^ ε ) | p ^ ε | 2 + 1 0 , 2 { τ ^ ε γ ( T s ^ ε ) 2 } 2 tr ( Y ^ ε ) + γ 2 ρ z * ( y ^ ε , s ^ ε ) | p ^ ε | 2 + ρ 0 .
Using tr ( X ^ ε + Y ^ ε ) 0 , derived from (79), now we reach
2 γ { 1 ( T t ^ ε ) 2 + 1 ( T s ^ ε ) 2 } + ( 1 w * ( x ^ ε , t ^ ε ) 1 ρ z * ( y ^ ε , s ^ ε ) ) ( γ 2 ) | p ^ ε | 2 2 tr ( X ^ ε + Y ^ ε ) + ρ 1 < 0 ,
which is a contradiction. □
The following corollary is a direct consequence of Theorem 5.
Corollary 3. 
Assume γ 2 , and let w and z be bounded viscosity solutions to (25) defined on Q T ¯ . Let
Q 0 = { ( x , t ) Q T w * ( x , t ) > 0 } { ( x , t ) Q T z * ( x , t ) > 0 } .
Then, if w * = w * = z * = z * on Q 0 \ ( Ω ¯ × { T } , it follows that w = z C ( Q 0 ¯ \ ( Ω ¯ × { T } ) ) .
An analogous result to Theorem 4 is the following theorem.
Theorem 6. 
Let w 0 , defined on Q T ¯ , be a bounded viscosity solution to (25) with γ 2 , and put
Q 0 = { ( x , t ) Q T w * ( x , t ) > 0 } , Q ˜ = Q T \ Q 0 .
Then, if
w * = w * = 0 o n   Q ˜ \ ( Ω ¯ × { T } ) ,
it follows that
w * = w * = 0 i n   Q ˜ .
Proof. 
Assuming
sup Q ˜ w * > 0 ,
we obtain 0 < γ 1 and ( x 0 , t 0 ) Q T \ Q 0 ¯ , satisfying
sup Q ˜ w * γ T t = w * ( x 0 , t 0 ) γ T t 0 > 0
by (83). Then, there is r 0 > 0 such that
w * ( x 0 , t 0 ) γ T t 0 w * ( x , t ) γ T t , ( x , t ) Q r 0 ( x 0 , t 0 ) .
Put
φ ( x , t ) = γ T t + w * ( x 0 , t 0 ) γ T t 0 ,
and see that that w * φ attains a local maximum 0 at ( x 0 , t 0 ) . Then, we obtain
0 ( φ 2 ) t + F ( φ , φ , 2 φ 2 ) = 2 w * ( x 0 , t 0 ) γ ( T t 0 ) 2 + w * ( x 0 , t 0 ) > 0 at   ( x 0 , t 0 ) ,
which is a contradiction.
It thus follows that w * 0 in Q ˜ , and hence in (84), from w 0 . □

5. Conclusions

In this paper, we study the quasilinear parabolic Equation (20) derived from the semilinear parabolic equation (1). Here, we review our results and discuss them in the relation to (1)–(3). To begin with, we note two facts regarding the relation between the exponents p and γ .
First, the condition p > 1 is equivalent to 4 < γ < in (22). The case γ = 4 arises if f ( u ) = e u in (2) by taking v = e 2 u . Several results of (20) obtained in this paper, however, are beyond this range of γ . The case γ < 2 means 0 < p < 1 for f ( u ) = u p , and then, v = 0 is equivalent to u = 0 . Hence, these results are associated with the quenching of the sub-linear parabolic equation. The other case of 2 < γ < 4 means p < 0 , and then v = 0 if and only if u = 0 . Thus, several properties of (20) presented in this paper are associated with the blow-up or quenching profiles of the solution to (1).
Second, Equation (1) for Ω = R N , N 2 , admits a singular stationary solution
u = C | x | 2 p 1
with a constant C > 0 if p > N N 2 . This exponent of p corresponds to 2 < γ < N + 2 , where v = v ( x ) 0 is realized by
v = C ( p 1 ) | x | 2 .
The role of the third exponent of this γ = N + 2 , other than γ = 2 and γ = 4 , however, has not been clarified yet.
Now we examine the results obtained in this paper in detail, in accordance with the theory of semilinear parabolic equations. Section 2 was devoted to the case of γ 2 , in which the convergence of a family of approximate solutions is discussed.
First, Theorems 1 and 2 are valid to 2 γ < 4 , with the non-trivial case 2 < γ < 4 corresponding to p < 0 . Since u 2 = 0 if and only if v = 0 in (11) for this p, this family of approximate solutions to (20) converges if a continuous quenching profile of the solution u = u ( x , t ) to (1) arises for p < 0 .
Second, Theorem 2 for γ = 4 and γ > 4 is associated with (2) for f ( u ) = e u and f ( u ) = u p with p > 1 , respectively. Since u 2 = + if and only if v = 0 in this case, a similar convergence is assured if there is a blow-up profile of the solution u = u ( x , t ) to (2) for these nonlinearities, which are continuous on Ω ¯ with the value + admitted.
Third, Theorem 6 is concerned with the case of γ 2 , with the non-trivial case γ < 2 corresponding to 0 < p < 1 . There, we obtain w v = 0 if and only if u = 0 , and therefore, this theorem is associated with the quenching profiles of the solution u = u ( x , t ) to (3) for f ( u ) = u p and 0 < p < 1 . In more detail, the quenching region
K = { x Ω u ( x , T ) = 0 }
determines the value of u = u ( x , T ) on the residual domain, Ω \ K .
The theorems obtained in this paper are thus strongly related to the results regarding the blow-up and quenching of the solution to semilinear equations. As we observe, several critical exponents on p are detected for these semilinear problems, which should lead to those of γ to (20) in accordance with the profile of v ( · , T ) .
We observe, also, that the theory of viscosity solutions guarantees the existence of the solution under the presence of the comparison theorem via the method of Perron [10]. There may be a challenge when taking this approach via Theorem 3 or Theorem 5, regarding their incompleteness, as compared to the standard case in which the positive region of the super-solution is not involved. We will come back to these problems in the future.

Author Contributions

Methodology, K.I., M.P. and T.S.; Writing—original draft, T.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors have no conflict of interest directly relevant to the content of this article.

References

  1. Masuda, K. Analysitc solutions of some nonlinear diffusion equations. Math. Z. 1984, 187, 61–73. [Google Scholar] [CrossRef]
  2. Baras, R.; Cohen, L. Complete blow-up after Tmax for the solution of a semilinear heat equation. J. Funct. Anal. 1987, 71, 141–174. [Google Scholar] [CrossRef]
  3. Lacey, A.A.; Tzanetis, D. Complete blow-up for a semilinear diffusion equation with a sufficiently large initial condition. IMA J. Appl. Math. 1988, 41, 207–215. [Google Scholar] [CrossRef]
  4. Galaktionov, V.A.; Vazquez, J.L. Continuation of blowup solutions of nonlinear heat equations in several space dimensions. Comm. Pure Appl. Math. 1997, 50, 1–67. [Google Scholar] [CrossRef]
  5. Ni, W.-M.; Sacks, P.E.; Tavantiz, J. On the asymptotic behavior of solutions of certain quasi-linear parabolic equations. J. Differ. Equ. 1984, 184, 61–73. [Google Scholar]
  6. Sakaguchi, S.; Suzuki, T. Interior imperfect ignition cannot occur on a set of positive measure. Arch. Ration. Mech. Anal. 1998, 142, 143–153. [Google Scholar] [CrossRef]
  7. Sakaguchi, S.; Suzuki, T. Nonexistence of solutions for a degenerate parabolic equation describing imperfect ignition. Nonlinear Anal. 1998, 31, 665–669. [Google Scholar] [CrossRef]
  8. Suzuki, T.; Takahashi, F. Capacity estimate for the blow-up set of parabolic equations. Math. Ann. 2008, 259, 867–878. [Google Scholar] [CrossRef]
  9. Lieberman, G.M. Second Order Parabolic Differential Equations; World Scientific: Singapore, 1996. [Google Scholar]
  10. Crandall, M.G.; Ishii, H.; Lions, P.-L. User’s guide to viscosity solutions of second order partial differential equations. Bull. Am. Math. Soc. 1992, 27, 1–67. [Google Scholar] [CrossRef]
  11. Koike, S. A Beginner’s Guide to the Theory of Viscosity Solutions; MSJ Memoirs Volume 13; Mathematical Society of Japan: Tokyo, Japan, 2004. [Google Scholar]
  12. Ladyžhenskaya, O.A.; Solonikov, V.A.; Uralćeva, N.N. Linear and Quasilinear Equations of Parabolic Type; AMS. Math. Soc. Transl. Math. Monograph: Province, RI, USA, 1968; Volume 23. [Google Scholar]
  13. Fellner, K.; Morgan, J.; Tang, B.Q. Global classical solutions to quadratic systems with mass control in arbitrary dimensions. Ann. I.H. Poincaré-AN 2020, 37, 281–317. [Google Scholar] [CrossRef]
  14. Baras, P.; Pierre, M. Problèmes paraboliques semi-linéaires avec donée measures. Appl. Anal. 1984, 18, 111–149. [Google Scholar] [CrossRef]
  15. Bothe, D.; Pierre, M. Quasi-steady-state approximation for a reaction-diffusion system with fast intermediate. J. Math. Anal. Appl. 2010, 368, 120–132. [Google Scholar] [CrossRef]
Table 1. Blow-up of the solution to semilinear parabolic equation.
Table 1. Blow-up of the solution to semilinear parabolic equation.
ReferencesProblemsMethods
[1,2,3,4]complete blow-upcomparison theorem
[9]post-blow-up continuationcomplex function theory
this paperblow-up pattenfunction analysis, theory of viscosity solutions
[5,6,7]estimates of the blow-up setmethod of isoperimetric inequality
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ishii, K.; Pierre, M.; Suzuki, T. Quasilinear Parabolic Equations Associated with Semilinear Parabolic Equations. Mathematics 2023, 11, 758. https://doi.org/10.3390/math11030758

AMA Style

Ishii K, Pierre M, Suzuki T. Quasilinear Parabolic Equations Associated with Semilinear Parabolic Equations. Mathematics. 2023; 11(3):758. https://doi.org/10.3390/math11030758

Chicago/Turabian Style

Ishii, Katsuyuki, Michel Pierre, and Takashi Suzuki. 2023. "Quasilinear Parabolic Equations Associated with Semilinear Parabolic Equations" Mathematics 11, no. 3: 758. https://doi.org/10.3390/math11030758

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop