Existence of solutions semilinear parabolic equations with singular initial data in the Heisenberg group
Abstract.
In this paper we obtain necessary conditions and sufficient conditions on the initial data for the solvability of fractional semilinear heat equations with power nonlinearities in the Heisenberg group . Using these conditions, we can prove that separates the ranges of exponents of nonlinearities for the global-in-time solvability of the Cauchy problem (so-called the Fujita-exponent), where is the homogeneous dimension of , and identify the optimal strength of the singularity of the initial data for the local-in-time solvability. Furthermore, our conditions lead sharp estimates of the life span of solutions with nonnegative initial data having a polynomial decay at the space infinity.
Key words and phrases:
Semilinear heat equation, Heisenberg group, Global existence, Lifespan estimates, Optimal singularities, fractional Laplacian2020 Mathematics Subject Classification:
Primary 35A01, Secondary 35K15, 35R03, 35R111. Introduction.
1.1. Heisenberg group .
This paper is concerned with nonnegative solutions of the fractional semilinear heat equation in the Heisenberg group
(1.1) |
with the initial data
(1.2) |
where , , , and is a nonnegative Radon measure on . Here, is the sub-Laplacian on the Heisenberg group and denotes the fractional power of on . For the exact definition of , see (1.6) below.
In this paper we show that every nonnegative solution of (1.1) has a unique nonnegative Radon measure in as the initial trace and study qualitative properties of it. Furthermore, we give sufficient conditions for the solvability of problem (1.1) with (1.2) and obtain sharp estimates of the life span of solutions with small initial data. Throughout of this paper, for and denote
The Heisenberg group is the Lie group equipped with the group law
and with the Haar measure on , where , , and is the scalar product in . It is well-known that the Haar measure on coincides with the -dimensional Lebesgue measure . The identity element for is and for all . The homogeneous Heisenberg norm is defined by
where is the Euclidean norm associated to . Then
is a left-invariant distance on . The homogeneous dimension of is . The left-invariant vector fields that span the Lie algebra are given by
for . The Heisenberg gradient is given by
and the sub-Laplacian is defined as
(1.3) |
where and stand for the Laplace operators on . In particular, the Heisenberg group is the most typical example of metric measure spaces with a structure different from that of .
1.2. Semilinear heat equations.
We recall the solvability of (1.1) in . For and , set . Let us consider nonnegative solutions of the fractional semilinear equation
(1.4) |
with the initial data
(1.5) |
where , , , and is a nonnegative Radon measure on . Here, denotes the fractional power of on .
Let us consider the case of first. The solvability of problem (1.4) with (1.5) has been studied in many papers since the pioneering work due to Fujita [F] (see, e.g., [QS19], which includes an extensive list of references for (1.4)). It is already known from his work and the subsequent works of Hayakawa [H73], Sugitani [S75], and Kobayashi–Sirao–Tanaka [KST77] that implies the nonexistence of any positive global-in-time solutions, while if , problem (1.4) with (1.5) possesses a positive global-in-time solution for appropriate initial data . For this reason, such an exponent is called the Fujita-exponent. Note that Sugitani [S75] also dealt with the fractional case .
Among these studies, Baras–Pierre [BP85] obtained necessary conditions for the solvability in the case of . Subsequently, the second author of this paper and Ishige [HI18] obtained a generalization to the fractional case and sufficient conditions for the solvability. Furthermore, they showed that every nonnegative solution of (1.4) has a unique nonnegatuve Radon measure on as the initial trace (see also [IKO20], which deals with the case where is a positive even integer as well as the case of ). More precisely, the following results have already been obtained.
-
(a)
Let , , and . Let be a nonnegative solution of (1.4) in , where . Then there exists a unique nonnegative Radon measure on such that
for all .
-
(b)
Let be as in assertion (a). Then must satisfy the following:
-
–
If , then ;
-
–
If , then for all ;
-
–
If , then for all .
Here is a constant depending only on , , and .
-
–
From this necessary condition (b) we can find a large constant with the following property:
- (c)
On the other hand, the following sufficient condition for the solvability has also been obtained:
- (d)
Note that these results show that is the Fujita-exponent. For (d), see also the papers written by Kozono–Yamazaki [KY94] , Robinson–Sierż\kega [RS13] (the case where and ), and Ishige–Kawakami–Okabe [IKO20] (the case where and ). The conditions (c) and (d) demonstrate that strength of the singularity at the origin of the function is the critical threshold for the local-in-time solvability of problem (1.4) with (1.5). We term such a singularity in the initial data an optimal singularity for the solvability of problem (1.4) with (1.5). For studies on optimal singularities, see e.g. [FHIL23, FHIL24, H24, HI18, HIT23, HS24, HT21, IKO20].
Let us go back to (1.1). For problem (1.1) with (1.2) in the case of , Zhang [Z98] proved that implies the nonexistence of any positive global-in-time solutions, while if , problem (1.1) with (1.2) possesses a positive global-in-time solution for appropriate initial data . Later, Pohozaev–Véron [PV00] considered more general equations in and proved that is included in the nonexistence case. Namely, we already know that is the Fujita exponent. Recently, Georgiev–Palmieri [GP21] obtained sharp estimates of the life span of solutions of problem (1.1) with small initial data in the case of and also proved the global-in-time solvability in the case of . Many other studies on the global-in-time solvability in have been undertaken, see e.g. [A01, AJS15, FRT24, JKS16, P98, P99, RY22]. We would like to emphasize that, in the case of , there were no previous studies on problem (1.1) with (1.2) to our knowledge.
In this paper we consider (1.1) and obtain analogous results to [HI18] (i.e. assertions (a)–(d)) in the Heisenberg group . Our conditions are optimal and, as an application, we show that our conditions can lead the sharp estimates of the life span of solutions with having a polynomial decay at the space infinity.
1.3. Notation and the definition of solutions.
In order to state our main results, we prepare some notation and formulate the definition of solutions. For any measurable set , denotes the Haar measure of . For any and , set . Furthermore, for any function , we set
Following [Folland], for we define the fractional sub-Laplacian by
(1.6) |
where is the Gamma function, is the heat flow, and is any function for which the relevant limit exists in norm. For the simplicity of notation, for , denote
Let be the fundamental solution of
(1.7) |
where . When , we briefly write instead of . For any locally integrable function in , it is well-known that
See e.g., [FS] (for ) and [MPS] (for ).
The key to our arguments is the following two sided estimate of the fundamental solution of (1.7).
Proposition 1.1.
satisfies the following estimate: there exist such that
(1.8) |
for and , where
For the proof of Proposition 1.1, see Subsection 2.3 below. In the case of , (1.8) is well-known (see e.g. [FS]). But in the case of , this is new. To prove the estimates (1.8) we employ a subordination formula in [G], which allows us to estimate the kernel via the heat kernel of the sub-Laplacian .
We formulate the definition of solutions.
Definition 1.1.
Let be a nonnegative measurable function in , where .
- (i)
- (ii)
- (iii)
1.4. Main results.
Now we are ready to state the main results of this paper. In the first theorem we show the existence and the uniqueness of the initial trace of solutions of (1.1) and obtain analogous results to assertions (a) and (b).
Theorem A.
Let , , and . Let be a solution of (1.1) in , where . Then there exists a unique Radon measure on such that
(1.11) |
for all . Furthermore, there exists depending only on , , and such that
-
(i)
if ;
-
(ii)
for all if ;
-
(iii)
for all if .
Since in (1.11) is unique, the following holds:
Remark 1.1.
Remark 1.2.
Remark 1.3.
As a corollary of Theorem A, we have:
Corollary 1.1.
Theorem A can be regarded as a generalization of the result in [HI18, Theorem 1.1]. Let be a solution of problem (1.1) with (1.2) in , where . We first prove the existence and the uniqueness of the initial trace of the solution and then obtain necessary conditions for the solvability (i.e. assertions (i)–(iii) in Theorem A). The proof of our necessary conditions in the case of follows the arguments in the proofs of [FHIL23, H24, HIT23, HS24, LS21], which enable us to more easily obtain the necessary conditions in [HI18, Theorem 1.1]. Let . Following these results, we get an integral inequality related to
(1.13) |
and then obtain the desired estimates by applying the existence theorem for ordinary differential equations to this estimate. See Lemma 2.10 below.
However, the proof in the case of is completely different from those in [FHIL23, H24, HIT23, HS24, LS21] and [HI18]. First, we extend solutions to the framework of weak ones and then employ a suitable cut-off function as a test function. This method was developed by Mitidieri–Pohozaev [MP01] and can be applied not only to semilinear heat equations but also to a wider class of equations. For the case of semilinear heat equations, see e.g. [GP21, IKO20, IS19]. In this paper we follow the arguments in [IKO20].
The reason why we adopt two methods comes from the fact that in Proposition 1.1, does generally not hold in the case of . In order to apply the arguments in [FHIL23, H24, HIT23, HS24, LS21], we have to get
for and . However, in the case of , we can not get such an estimate from Proposition 1.1, since is an exponential function. On the other hand, in the case of , it can be regarded as and we can avoid this problem, since is a polynomial function. In the previous studies dealing with [H24, HI18, HIT23, HS24, LS21, FHIL23], this problem did not arise because they were either dealing with the fractional case or had an explicit formula of .
By Theorem A we have
Theorem B.
We give sufficient conditions for the solvability of problem (1.1) with (1.2). The proofs follow the arguments in [HI18].
Theorem C.
Theorem D.
Theorem E.
Corollary 1.2.
Let , , and . Define
Assume that in for some and . Then there exists with the following properties:
- (1)
- (2)
In particular, if and , the solution in the assertion (1) is a global-in-time one.
From this corollary it can be seen that is an optimal singularity for the local-in-time solvability. In particular, together with Remark 1.3, it can be seen that is the Fujita-exponent.
The rest of this paper is organized as follows. In Section 2 we collect properties of and and prepare some preliminary lemmas. In Section 3 we prove (1.11) in Theorem A and Theorem B. In Section 4 we prove assertions (i)–(iii) in Theorem A and complete the proof of Theorem A. In Section 5 we prove Theorems C–E. In Section 6, as an application of our theorems, we obtain estimates of the life span of solutions of problem (1.1) with small initial data.
2. Preliminaries.
In what follows, the letters and denote generic positive constants depending only on , , and . For any two nonnegative functions and defined on a subset , we write for all if for all , and we write for all if and for all . Furthermore, for , we write if for some constants .
2.1. Basic properties of and .
In this subsection we collect properties of the Heisenberg group and the fundamental solution . The following lemma is used when we calculate integrals in the Heisenberg group and is well-known (see e.g. [BHQ24, P98]). Therefore, we omit the proof.
Lemma 2.1.
Let , , , be a continuous function. Then one has
(2.1) |
Moreover, one has
(2.2) |
For and , define
The following properties are taken from [FS, MPS]. Given , the function has the following properties:
(2.3) | |||
(2.4) | |||
(2.5) | |||
(2.6) |
for all , , and .
Since , for and , by integration by parts we have:
(2.7) |
for all .
2.2. A covering lemma and the Hardy-Littlewood maximal function.
For and a nonempty subset , set
Lemma 2.2.
-
(a)
Let and . Then there exists an universal constant such that we can find a family of balls for some countable family of indices such that
-
(i)
;
-
(ii)
.
-
(i)
-
(b)
Let and . Then for each , we can find a family of balls for some countable family of indices such that
-
(i)
;
-
(ii)
for each and ;
-
(iii)
, where is a constant independent of , and .
-
(i)
Here, for any set , denotes the cardinal number of .
Proof.
Since the proof of (a) is similar to that of (b) and even easier, we need only to prove (b). Fix . We consider the following family of balls , which covers . By Vitali’s covering lemma, we can extract a disjoint family of balls for some countable family of indices satisfying
By the construction, it is straightforward that the family satisfies (i) and (ii). In addition, by (2.2),
which implies (iii). This competes our proof. ∎
Recall that the Hardy-Littlewood maximal function is defined by
where the supremum is taken over all balls containing and is the radius of . It is well-known that is bounded on for .
We have the following result whose proof is quite elementary and will be omitted.
Lemma 2.3.
For , there exists such that
for all , and .
2.3. Some kernel estimates.
In this subsection, we obtain estimates of the fundamental solution and collect the basic properties of .
Proof of Proposition 1.1.
The case is well-known. In fact, we have
(2.8) |
for all and . See e.g. [FS].
It remains to prove for the case . We use the following subordination formula in [G]:
(2.9) |
where the function satisfies the following properties:
-
(i)
and ;
-
(ii)
;
-
(iii)
for all ;
-
(iv)
for all ;
-
(v)
for all .
We first establish an upper bound for . By (2.8), (2.9), (iii), and (iv), we have
Case 1: . Using the change of variable ,
Case 2: . In this case, by (2.9), (2.8), and (ii),
where in the second line we used the change of variable and in the last inequality we used (v). We have proved that
for and , which is the upper bound in (1.8).
Lemma 2.4.
For each and for any there exist such that
for and . Consequently, for we have
for all and .
Proof.
Lemma 2.5.
Let . For every , we have
for all .
Proof.
Lemma 2.6.
Let and . Then for every .
Proof.
Lemma 2.7.
Let and . Then we have
-
(a)
We have
for .
-
(b)
Let . Then
for .
Proof.
Since the proof of (a) is similar to (b) (even easier), we only give the proof of (b).
For any Radon measure on , we define
for and .
We have the following estimate.
Lemma 2.8.
Let and . For any Radon measure on , there exists a constant such that
for .
2.4. Preliminary lemmas.
At the end of Section 2, we provide some lemmas to prove the solvability.
Lemma 2.9.
Proof.
The proof is quite standard. See e.g. [HI18, HIT23, IKS16, RS13]. However, we would like to provide it for the sake of completeness. Define as follows. Set and define
(2.10) |
for . Let be a supersolution of problem (1.1) with (1.2) in , where . Then it follows inductively that
for a.a. and . It follows that
for a.a. and . This, along with (2.10), implies that is a solution of problem (1.1) with (1.2). Since any solution is also a supersolution, if is another solution of problem (1.1) with (1.2) then a similar argument also shows that . Hence, is a minimal solution. This completes our proof. ∎
The key to the proof of Theorem A in the case of is the following lemma on the existence of solutions of ordinary differential equations. This idea comes from [LS21].
Lemma 2.10.
Let be a nonnegative measurable function on for some . Assume that
(2.11) |
where , , , and . Then there exists such that
In addition, if , then
3. Initial trace.
In this section we show the existence and uniqueness of the initial trace and prove (1.11). The proof follows the arguments in [HI18].
Lemma 3.1.
Let be a solution of (1.1) in , where . Then
(3.1) |
for all and . Furthermore, there exists a unique Radon measure on such that
(3.2) |
for all .
Proof.
Let . Then, from (1.9) and the nonnegativity of , there exists such that
for a.a. and . It follows that
From (1.8), we have
which implies
for a.a. . It follows (3.1) since .
We now take care of (3.2). It suffices to prove (3.2) for . From (3.1), the Riesz representation theorem [Cohn, Theorem 7.2.8] and the weak compactness of Radon measures [Simon, 4.4 THEOREM], we can find a sequence with and a nonnegative Radon measure on such that
(3.3) |
for all .
For the uniqueness of the Radon measure , we assume that there exist a sequence with and a nonnegative Radon measure on such that
(3.4) |
for all . Taking a subsequence if necessary, we might assume that for . From (1.9),
for a.a. . From the above inequality, the fact , and (2.3), we have
Letting , by Lemma 2.7, (3.1), (3.3), and (3.4) we obtain
for all . Similarly, it follows that
for all . This completes our proof. ∎
Lemma 3.2.
Proof.
Without loss of generality we need only to verify (3.5) with . Let such that . By Lemma 3.1, we can find a unique Radon measure on such that
(3.6) |
for all . Furthermore, from (1.9) we have
(3.7) | ||||
for a.a. and . On the other hand, since , by (2.3) we have
(3.8) | ||||
for a.a. and . This, in combination with (3.6) and Lemma 2.7, yields
for some and a.a. . Taking this, (3.7) and (3.8) into account, we obtain
for a.a. . By invoking (3.6) and Lemma 2.7,
This, together with (1.10), implies that
Hence, it suffices to prove the following
We consider two cases: and .
Case 1: . In this case, by using (1.8),
for and . This, together with the fact (see Lemma 2.5), implies
for all and .
On the other hand, for and , by (1.8),
Therefore, for , by Lemma 3.1 we have
This, in combination with Lemma 2.7 and the Lebesgue Domination theorem, implies that
as desired and hence this completes the proof for the case .
Case 2: . The proof of this case has the same spirit as that of the case . Since is bounded on , it follows from (1.8) that,
for and . This together with the fact (see Lemma 2.5), implies
for all and . On the other hand, for and , by (1.8),
Therefore, for , by Lemma 3.1 we have
This, in combination with Lemma 2.7 and the Lebesgue Domination theorem, implies that
This completes the proof for the case . The proof is complete. ∎
4. Necessary conditions for the solvability.
In this section we prove assertions (i)–(iii) in Theorem A and complete the proof of Theorem A. Furthermore, we also prove Theorem B. The proof of Theorem A is quite long and will be divided into two cases: and .
The case
Lemma 4.1.
Let , , and . Let be a solution of (1.1) in , where . Then there exist positive constants and depending only on , and such that
for all with and for a.a. .
Proof.
Let be sufficiently small. Let
(4.1) |
We set for a.a. and . Then it follows from (1.9) that
(4.2) |
for a.a. and . We shall show that
(4.3) |
for a.a. and . From (1.8),
for all . Then, using (4.2) with , we have
for a.a. with and . It follows (4.3). Furthermore, by (1.9), (2.5), and (4.2),
which implies for ,
for a.a. and . From (1.8),
and
for . This, in combination with (2.6) and Jensen’s inequality, further implies
for a.a. and .
We now set
Then we can rewrite the above inequality as
for a.a. .
In the case of , note that . It follows from Lemma 2.10 with that
for all and a.a. . In the case of , similarly we have
for all and a.a. . This completes our proof. ∎
Proof of Theorem A in the case of .
The case In this case, we consider solutions of problem (1.1) with (1.2) in the following weak framework.
Definition 4.1.
Proof.
Proof of Theorem A in the case of .
The proof follows the arguments in [IKO20, Theorem 1.2]. Let be a solution of problem (1.1) with (1.2) in , where . From the proof in the case of , it is sufficient to show that initial data satisfies assertions (i)–(iii) in Theorem A. It follows from Lemma 4.2 that satisfies (4.4). Let . In what follows, denote
Note that
Let
Set
Then and
Set
Since , for any , we can find such that
(4.5) |
For any , we set
and for any , we set
For the simplicity of notation, set
Note that
(4.6) |
for all and . Then we shall calculate the derivatives of . Since , by (4.5) we may assume that there exists such that and satisfy
(4.7) |
(4.8) |
By (1.3), (4.6), and (4.8) we see that
(4.9) |
Substituting into (4.4), by (4.9) and Hölder’s inequality, we obtain
(4.10) |
for all . On the other hand, it follows from (2.2) that
for all . This, together with (4.10), implies that
(4.11) |
for all . Let be a sufficiently small positive constant. For any , set
(4.12) |
Since is decreasing on and , for any with , we have
(4.13) |
Since if , by (4.12) and (4.13) we obtain
(4.14) |
Therefore, we deduce from (4.11), (4.12), and (4.14) that
for all . Therefore, we have
(4.15) |
Since
by (4.15) we obtain
Letting , we see that
(4.16) |
for all .
Setting , similarly to the proof in the case of , we obtain
for all . Thus, assertions (i)–(iii) in Theorem A follow and the proof is complete. ∎
Proof of Corollary 1.1.
At the end of this section, we give a proof of Theorem B.
Proof of Theorem B.
Fix . Let be a solution of (1.1) in , where . Let and . For each , we have
By the covering lemma in Lemma 2.2 a family of balls for some index set such that
-
(i)
;
-
(ii)
for each ;
-
(iii)
for each and some universal constant .
This, together with Theorem A and (1.8), implies that
(4.17) | ||||
Similarly, we have
(4.18) | ||||
From these two above estimates (4.17) and (4.18), Theorem A and Lemma 3.1 we obtain
(4.19) |
for a.a. . Let be such that in , on , outside . It follows from (4.19) that
(4.20) | ||||
for and a.a. . By Lemma 3.2 and the fact that is continuous, we see that
(4.21) |
Furthermore, by Lemmas 3.1and 2.4 we have
(4.22) | ||||
From (4.20), (4.21), and (4.22),
for . This, in combination with (4.17) and (4.18), implies that
This and (1.10) yield that is a solution of problem (1.1) with (1.2) in . This completes our proof. ∎
5. Sufficient conditions for the solvability.
Proof of Theorem C.
It suffices to consider the case of . Indeed, for any solution of problem (1.1) with (1.2) in , we see from (2.4) that with is a solution of problem problem (1.1) with (1.2) in . Set . Then it follows from (2.5) and Lemma 2.8 that
This, together with the assumption of Theorem C and , implies that
for all . Therefore, taking a sufficiently small if necessary, we obtain for . This means that is a supersolution of problem (1.1) with (1.2) in . Then the theorem follows from Lemma 2.9. ∎
Proof of Theorem D.
For the same reasons as in the proof of Theorem C, it suffices to consider the case of . Assume (1.14). We can assume, without loss of generality, that . Indeed, if , then, for any , we apply Jensen’s inequality to obtain
for all . Thus, (1.14) holds with replaced by .
Let and set . It follows from (2.5), (2.6) and Jensen’s inequality that
for all . Furthermore, by Lemma 2.8 and (1.14) we have
This implies that
Therefore, taking a sufficiently small if necessary, we obtain for . At this stage, arguing similarly to the proof of Theorem C we complete the proof of Theorem D. ∎
Proof of Theorem E.
For the same reasons as in the proof of Theorem C, it suffices to consider the case of . Let and be as in (1.15). Let be such that
-
(a)
is positive and convex in ;
-
(b)
and are monotone increasing in .
Let be a nonnegative measurable function in satisfying (1.16). Simce for , it follows that
(5.1) |
for all and . Set
By (5.1) we apply Lemma 2.8 to obtain
which implies that
(5.2) |
for . Define
Then by (2.5), (2.6), and Jensen’s inequality,
(5.3) |
for . On the other hand, by property (b) and (5.2) we see that
(5.4) |
for . By (1.15) we have
for all . Since for all , it follows that
for all . These together with (5.4) imply that
(5.5) |
for all . Similarly, by (5.2) and property (b) we have
(5.6) |
for all . By (5.5) and (5.6) we obtain
(5.7) |
for all . Therefore, taking a sufficiently small if necessary, we deduce from (5.3) and (5.7) that for . At this stage, arguing similarly to the proof of Theorem D we complete the proof of Theorem E. ∎
Finally, we give a proof of Corollary 1.2.
Proof of Corollary 1.2.
Assume that for some and . Let and fix so that it satisfies
It follows from (2.1) and (2.2) that
for . Then taking sufficiently small and if necessary, we see that (1.14) holds for all . This implies that problem (1.1) with (1.2) possesses a solution in . If , we can take . Let and . By similar calculations, we see that
for all , where and is as in (1.15). Then taking sufficiently small and if necessary, we see that (1.16) holds for all . This implies that problem (1.1) with (1.2) possesses a solution in .
6. Application.
Since the minimal solution is unique, we can define the life span as the maximal existence time of the minimal solution of problem (1.1) with (1.2).
For (1.4) and in the case of , Lee–Ni [LN92] obtained sharp estimates of as by use of the behavior of at the space infinity. Subsequently, the second author of this paper and Ishige [HI18] obtained a generalization to the case of . Recently, Georgiev–Palmieri [GP21] obtained a generalization of [LN92] to the Heisenberg group in the case of . In some cases, however, sharp estimates have not yet been obtained.
In this section, as an application of our theorems, we show that similar estimates of as in [LN92, HI18] in the Heisenberg group . Theorems F and G are generalizations of [LN92, Theorem 3.15 and Theorem 3.21], respectively. At the end of this section, summaries of these theorems and previous study [GP21] are given.
Theorem F.
Let and . Let and be a nonnegative measurable function in such that
for a.a. .
-
(i)
Let and . Then there exists a constant such that
for all sufficiently small .
-
(ii)
Let or . Then there exists a constant such that
for all sufficiently small .
For the simplicity of notation, we denote . We give a proof of Theorem F.
Proof of Theorem F.
Since , by Theorem D we have
for all sufficiently small . This implies that . So, we can assume without loss of generality that is sufficiently large.
We apply Theorem A to prove Theorem F and assume that problem (1.1) with possesses a solution in . For any , we see that
(6.1) |
for all and sufficiently small . In the case of , it follows from assertion (ii) in Theorem A that
for all and sufficiently small . This implies that
(6.2) |
(6.3) |
for all sufficiently small . By (6.1) and (6.2) with we obtain assertion (i). Furthermore, by (6.1) and (6.3) with we obtain assertion (ii) in the case where and .
Theorem G.
Let and . Let and be a nonnegative measurable function in such that
for a.a. .
-
(i)
Let and . Then there exists a constant such that
for all sufficiently small .
-
(ii)
Let or . Then there exists a positive constant such that
for all sufficiently small .
Proof.
We apply Theorem E to prove assertion (i). Let and set
for (see (1.15)). For any and , set
We shall prove that problem (1.1) with possesses a solution in . Let be such that
Then we see that and for all .
We consider the case of . By (2.1) and (2.2), we have
(6.5) |
for all . Since
(6.6) |
for sufficiently small , we have
(6.7) |
for all and sufficiently small . This together with (6.5) implies that
for all and sufficiently small . Therefore, taking a sufficiently small if necessary, we apply Theorem E to see that problem (1.1) with possesses a solution in and
for all sufficiently small .
We consider the case of . Similarly to (6.6) and (6.7), we have
for all and sufficiently small . Then we apply the same argument as in the case of to see that
for all sufficiently small . Thus, assertion (i) follows.
We shall prove assertion (ii) in the case where and . It follows that . For and , set
Let be such that . Then by (2.1) and (2.2) we have
(6.8) |
for all and . On the other hand, it follows that
(6.9) |
for all . Taking a sufficiently small if necessary, by (6.8) and (6.9) we obtain (1.14) for all . Then it follows from Theorem D that problem (1.1) with possesses a solution in and
Thus, assertion (ii) in the case where and follows.
It remains to prove assertion (ii) in the case of . For and , set
It follows from (2.1) that
Then taking a sufficiently small if necessary, we obtain
for all sufficiently small , where is as in Theorem C. Then it follows from Theorem C that problem (1.1) with possesses a solution in and
Thus, the proof is complete. ∎
At the end of this section, we describe summaries of Theorems F and G in tables. The following tables show the behavior of the life span as , where
If it is marked with †, it is already shown in [GP21] in the case of .
† | |||
(, see Table 2) | † | ||
† |
Acknowledgments.
The first-named author was supported by the research grant ARC DP220100285 from the Australian Research Council.
Statements and Declarations.
-
•
The second-named author did not receive support from any organization for the submitted work.
-
•
The authors have no relevant financial or non-financial interests to disclose.