Global boundedness and blow-up in a repulsive chemotaxis-consumption system in higher dimensions
Abstract
This paper investigates the repulsive chemotaxis-consumption model
in an -dimensional ball, , where the diffusion coefficient is an appropriate extension of the function for some . Under the boundary conditions
we first demonstrate that for , or with , the system admits globally defined classical solutions that are uniformly bounded in time for any choice of sufficiently smooth radial initial data. This result is further extended to the case when is chosen to be sufficiently small, depending on the initial conditions. In contrast, it is shown that for , the system exhibits blow-up behavior for sufficiently large .
keywords:
repulsive chemotaxis-consumption system, blow-up, global boundednessMSC:
[2010] 35B44 , 35K51 , 92C171 Introduction
Since the pioneering work of Keller and Segel in the early 1970s [10, 11], the mathematical investigation of chemotaxis – the directed movement of organisms in response to chemical gradients – has significantly enhanced the understanding of various biological mechanisms. Of particular interest is the chemotactic behavior of aerobic bacteria that consume oxygen as a nutrient, a subject that has attracted considerable attention due to its complex dynamics observed in both experimental and simulation settings [6, 20]. Accordingly, the following equations have been derived and subjected to rigorous analysis as a fundamental model of this nutrient-based chemotaxis interaction [15]:
(1) |
where denotes the density of the organisms and stands for the concentration of nutrients, such as oxygen.
When , , and are positive and appropriately chosen, it is well-known that the Neumann problem for (1) possesses an energy dissipation structure, arising from the interplay between the consumption term and the chemotaxis term , of the form in its simplest prototype case where , , and :
(2) |
The energy dissipation structure (2) has played a crucial role in demonstrating global-in-time boundedness, stability, and asymptotic behaviors of solutions as well as global solvability (c.f. [15]).
Specifically, beyond the small initial data case where the global existence and asymptotic behaviors converging to a constant steady state have been proved using weighted -estimates [16, 31], global large-data classical and weak solutions have been achieved via exploiting (2) in the two and three dimensions, respectively [25] (see [9, 18] for eventual smoothness and asymptotic behaviors of such weak solutions in three dimensions).
Moreover, variants of the dissipation structure (2), capable of allowing singular behaviors of , have been employed to see the global solvability and asymptotic behaviors of solutions. Indeed, the global existence of the classical solution has been proved for general and involving and with for [2]. Later, this result has been improved to for [12]. For and , weak solutions have been constructed when the initial data is sufficiently small [21]. In the case of , another dissipation structure that provides weak regularity than (2) has been employed to prove global solvability of generalized solutions for [26] (c.f. [14]), and of renormalized solutions for [27].
Furthermore, similar quantitative results have been considered when is of porous medium type, , or its nondegenerate form, with , by using this enhancing diffusivity or variants of (2). Exploiting the diffusivity of , in presence of general and , the existence of global bounded solutions has been shown for and in a weak sense [19], and for and in the classical sense [23]. In the case of , the global solvability has been shown for and using a variant of (2) [22]. Later, this result has been improved by [7], where the global well-posedness and asymptotic behaviors are established for and [7]. In the case of , the global solvability has been proven for and [13], and for and in a generalized concept [29]. The latter has been extended to the tensor-valued satisfying with for [28].
For parabolic-elliptic simplification of (1) with , , and , the global well-posedness and asymptotic behavior have been shown under Robin boundary condition by Fuest-Lankeit-Mizukami [8] when (see also [1]). We refer to [30] for its Dirichlet counterpart.
Remarkably, the parabolic-elliptic counterpart of (1) may exhibit blow-up behavior when is of repulsion type, .
Indeed, under No-flux/Dirichlet boundary conditions, finite-time blow-up solutions have been found for , , and by Wang-Winkler [24] when . This result has also been verified for and [3]. Additionally, in the latter case, the global existence of bounded solutions has been established when has a positive lower bound. However, to the best of our knowledge, there are no results concerning the existence or blow-up of solutions for and .
Our main objective in this paper is to either establish the global existence of bounded solutions or construct unbounded solutions for the repulsion-type chemotaxis-consumption system in dimensions three and higher. To be more precise, we consider the following initial-boundary value problem
(3) |
In our analysis, the spatial domain is specifically chosen as , a ball of radius centered at the origin in . The diffusion coefficient is chosen to extend the prototype choice as follows:
Throughout this paper, the initial function is assumed to satisfy
(4) |
Before presenting the main results, we begin by outlining the local existence and blow-up criteria for (3). Although these topics have been widely explored in the literature for more generalized frameworks [5], this paper specifically addresses our particular case.
Proposition 1.
Let , with , and . Assume that
(5) |
and that the initial function satisfies (4). Then, there exists a maximal time for which the problem (3) possesses a unique classical solution in , which is radially symmetric and positive over , with the following regularity properties:
Furthermore, if , then blows up in the sense that
Additionally, the mass of is conserved over time:
(6) |
and is uniformly bounded in time:
(7) |
The first objective of this study is to investigate conditions on the diffusion rate that ensure the existence of globally bounded regular solutions. The following theorem demonstrates that, within this radially symmetric setting, the existence of a global-in-time bounded solution is assured for certain values of , specifically or with a constraint on .
Theorem 1.
A subsequent natural question is under what conditions the boundedness or unboundedness of the solution is determined for the case . It might be expected to experts that small data implies the existence of global bounded solutions. However, we can not find it in the literature and therefore, for clarity, we first demonstrate that bounded solutions can be constructed globally in time when the Dirichlet data is sufficiently small.
Theorem 2.
Finally, we establish that when is sufficiently large, blow-up of the solution may occur, in particular, in a more restricted range .
Theorem 3.
Remark 1.
We remind that in two dimensions, blow-up solutions have been constructed for any [3]. We may also expect that the blow-up could occur in the range for as well. We leave it as an open question.
Plan of the paper Theorem 1 extends the result in [3] to higher dimensions. Whereas the authors in [3] exploited the uniform smallness of near the origin to ensure uniform boundedness of , thereby achieving global boundedness of for , our proof begins with a pointwise estimate of near the origin (Lemma 2). This estimate is derived from the uniform boundedness of and the assumption of radial symmetry. Employing this approach, we can control the mass accumulation function near the origin through a comparison method for the equation of (Lemma 4). This control leads to an improved regularity of (Lemma 5) near the origin, which, when combined with estimates away from the origin (Lemma 1), ensures global boundedness of through -estimates in higher dimensions (Lemma 6).
For Theorem 2, we first exploit the smallness assumption on , depending on , to obtain the behavior of via a comparison argument similar to that used in Lemma 4 (Lemma 7). This behavior of , in conjunction with a Hardy-type inequality (Lemma 8) and a one-dimensional variant of the Gagliardo-Nirenberg inequality, allows us to derive -estimates for (Lemma 9).
In Theorem 3, the method follows the approach employed in [3] (see also [24]).
This method involves tracking the time evolution of the quantity with appropriately chosen .
Obtaining a suitable lower bound for is crucial in this process. The authors in [3] provided this lower bound by appropriately estimating and , respectively.
However, while the lower bound for can be efficiently estimated even in higher dimensions using a variant of the ODE comparison method (Lemma 10, and see [24, Lemma 3.1] for the proof), demonstrating the polynomial decay of , which was straightforward in dimensions two [3], seems difficult to demonstrate for dimensions three and higher. As a result, it is unlikely to obtain as in [3].
To overcome the difficulty,
instead, we make full use of the second equation of (3) and the estimate of at the spatial origin (Lemma 11) to obtain an appropriate ordinary differential inequality (ODI) for (Lemma 12), which is a technical novelty compare to [3].
It turns out that such ODI satisfies , namely, has
a linear growth in time, which implies blow-up in finite or infinite time. On the other hand, in three dimensions, since it is shown that is uniformly bounded, blow-up must occur in a finite time.
For clarity throughout this paper, we abbreviate as when no confusion arises. Constants are denoted by with , where the subscript serves to distinguish these constants within each chapter. Additionally, indicates a dependency of the constant on variables , and others. Finally, we denote by the measure of the dimensional unit sphere .
2 Global boundedness for
Hereafter, we fix , with , and assume that satisfies (5). we denote by the solution in as in Proposition 1, and denote
Utilizing the radial symmetry, the equations in (3) can be rewritten as the following scalar equations with a slight abuse of notation , where and :
(11) |
(12) |
Hereafter, we fix , with , and assume that satisfies (5).
2.1 Estimates of
In the subsequent two lemmas, pointwise estimates for are presented. The estimates are two-fold: (Lemma 1) away from the spatial origin and (Lemma 2) near the origin. The first estimate indicates that the origin is the only possible blow-up point for , while the second leads to Lemma 4, the behavior of near the origin.
We first observe that is positive and bounded away from the origin.
Lemma 1.
Let be a solution given in Proposition 1. It holds that is positive in . Moreover, for any choice of , we have
Proof.
The following lemma shows that near the origin uniformly in time. Our basic approach involves employing a Newtonian kernel-type test function .
Lemma 2.
Let be a solution given in Proposition 1. For any , it follows that
Proof.
In light of the positivity of and , it holds through integration by parts and the positivity of that
for all . Let be arbitrarily chosen. Since implies and is positive as shown in Lemma 1, we obtain the desired result. ∎
2.2 On the mass accumulation function of
We now examine the mass accumulation function of given by
(13) |
for . Clearly, is nonnegative and nondecreasing. Based on the observation that blow-up of the solution can only occur near the origin (Lemma 1), our analysis primarily focuses on the behavior of near the origin. As a preparation, the following simple calculation is needed.
Lemma 3.
Proof.
In the next lemma, we show that when , the inequality (14) provides the behavior of near the origin.
Lemma 4.
Proof.
Given the condition , we first fix , choose , and define as
Let . By performing a fundamental calculation on , we obtain
In addition, we note that along the parabolic boundary of . Indeed, it is trivial that and for all . Furthermore, at , we find that
for all . Here, the last inequality of the above holds due to the fact that . We now define an auxiliary function over the domain . Then, we have
To prove the inequality in (16), we employ a variant of the comparison principle: Let and define the set
Assume, for the sake of contradiction, that . Let Then, due to on the parabolic boundary of . Thus we can choose fulfilling . Note that , , and . Therefore, it holds that
which contradicts . This implies that . Since can be taken arbitrarily small, it follows that for every , thereby proving (16). ∎
2.3 Proof of Theorem 1
The upper estimate of established in Lemma 4 leads to a more refined behavior for near the origin compared to that in Lemma 2.
Lemma 5.
Proof.
Note that it is enough to consider the case . Let . We fix and as in Lemma 4, and then define
and
Since the definition of yields , it holds that
that is, . For fixed , it follows from (12) and (7) that
(19) |
which, along with (16), proves (17). Furthermore, (17) leads to (18),
This completes the proof. ∎
We are ready to prove the boundedness of in for large .
Lemma 6.
Proof.
Let us fix . We first consider the case that
Let , where is the constant appeared in Sobolev embedding theorem
For such , we take as in Lemma 5. By applying Lemma 1 with , we obtain that is bounded in . Thus, it holds that
(21) |
with some . For fixed , we now observe standard -estimates for :
By Sobolev embedding, the Gagliardo-Nirenberg inequality, and Young’s inequality, we have
where , , and . Hence, it follows from (21) and Young’s inequality that for some
Similarly,
where and . Therefore, it follows from the above estimates that
(22) |
Next, in the case of
we use Lemma 1 and Lemma 2 to control . In a similar way, let , and we see that
(23) |
We first note that
(24) |
On the one hand, a simple calculation and Hölder’s inequality yields
where and is such that By Sobolev embedding, it follows that
where , , and . From Lemma 1 and Lemma 2, we can show that is bounded in , which induces that is bounded in due to the fact that
Thus, noticing that , we obtain through Young’s inequality that
(25) |
with some . Hence, combining (24) and (25) into (23), we see that
(26) |
In addition, the Gagliardo-Nirenberg inequality, Young’s inequality, and (6) yield that for any and , we can find such that
Therefore, writing , both (22) and (26) can be rewritten by
which induces (20). ∎
Proof of Theorem 1.
Note that the standard elliptic regularity theory and Lemma 6 with show is bounded uniformly in time. Thus, as in (23) and (24), we have for some
for all . Now, employing a standard iteration method (see e.g., [4, 17]), we can show that is uniformly bounded in , which proves Theorem 1 with the help of blow-up criteria in Proposition 1. ∎
3 Global boundedness for
Lemma 7.
Proof.
We begin by noting that is finite since satisfies (4). For fixed , since (15) and (19) are still valid, we have
We again use the comparison principle, as in the proof of Lemma 4; we first observe that
Thus, for any
we have Furthermore, it follows from the definition of that for all and ,
Consequently, as in the proof of Lemma 4, we can conclude (27). ∎
We now proceed to the standard - estimates for . As a preliminary step, we introduce a simplified version of the Hardy inequality on balls.
Lemma 8.
For any radially symmetric function , , there exists such that
(28) |
Proof.
By applying integration by parts and Young’s inequality, we can derive
(29) | ||||
with some . Note that it is enough to consider the case . By utilizing the standard trace embedding and subsequently applying the Gagliardo-Nirenberg inequality along with Young’s inequality, we obtain
(30) | ||||
where , , and . In consequence, by employing (29) and (30) together, we have (28). ∎
Lemma 9.
Proof.
For given , we pick such that
(32) |
Then, we choose to satisfy
(33) |
For fixed , we multiply the first equation in (3) by and integrate by parts, then add to both sides to obtain
(34) | ||||
Define a nonnegative function Given the nonnegativity of and the increasing nature of the function , it follows from integration by parts again, uniform boundedness of as established in (7), and the Sobolev embedding that
(35) | ||||
For convenience, we define and . By combining (34) and (35), and then applying Young’s inequality and Hölder inequality, we obtain
(36) | ||||
with some , where is given by (33). Since
due to (33), it follows that
(37) |
where . Next, using a one-dimensional version of the Gagliardo-Nirenberg inequality, we observe that
(38) |
with some , where
We note that
due to (33). In addition, we have
which, with an application of Lemma 8 with and , entails that
(39) |
with some . Moreover, we use integration by parts to estimate
(40) | ||||
where . Since by (32), applying (27) to (40) allows us to deduce for some
(41) |
where is a number defined in Lemma 7. Therefore, by combining (39) and (41) with (38), we infer that
(42) |
where . Consequently, by plugging (37) and (42) into (36), we can choose using Young’s inequality that
Therefore, the function satisfies
thereby leading to the result in (31). ∎
Proof of Theorem 2.
Let satisfy (32). By Lemma 9, there exists such that for all . Let be such that . For fixed , we employ again (34) and (35) to obtain, for some ,
(43) |
By the Gagliardo-Nirenberg inequality, we see that for some
(44) | ||||
Here, due to the condition on , we have Furthermore, since fulfills , it follows that Thus, substituting (44) into (43) and applying Young’s inequality, we establish the -boundedness of for any . Hence, similar to the proof of Theorem 1, we can apply the standard iteration method to achieve the uniform boundedness of in . ∎
4 Blow-up in a finite or infinite time for
This section aims to detect the blow-up phenomena that occur due to the competition of diffusive and cross-diffusive dynamics. This will be verified through the analysis of a temporal evolution involving mass accumulation function .
4.1 Estimates for at the origin
We begin by addressing the lower bound for . For a detailed proof, we refer to [24, Lemma 3.1].
As a result of Lemma 10, the value of at the origin can be controlled by with a quantity involving .
Lemma 11.
4.2 ODI for a functional involving
We derive an ODI for , which is crucial for our blow-up argument.
Lemma 12.
Proof.
Let . Obviously, is well-defined for such , and the regularity property of is guaranteed via the dominated convergence theorem. In order to see (47), we first define
(48) |
We recall that satisfies the following equation (15) in :
Then, is formulated as the following:
From the assumption (10) and (48), we have
which entails, via integration by parts, that
(49) |
where we used the condition to see that
Applying Hölder inequality and integration by parts, we further compute the rightmost term of (49) as
(50) |
Here, the first integral on the right-hand side is finite because implies
Hence, in light of (49) and (50), one can find such that
In order to estimate , we first observe from (12) that
(51) |
for all . Using (51), , and integration by parts, we find
(52) |
We employ (12) again to see that
which yields, due to the nonnegativity of the last term in (52) and (7), that
(53) |
We observe from Hölder inequality that
(54) |
where the last integral is finite due to the condition . Therefore, combining (53) and (54), it follows that
with some . Hence, we can conclude (47). ∎
We are ready to present the proof of Theorem 3.
Proof of Theorem 3..
Let , and fix as in Lemma 12. Plugging (46) into (47) shows that
(55) |
for all , where and are constants specified in Lemma 12. For and , let us define
and fix a constant depending
We choose such that
Since is increasing, it holds that for and
(56) |
Upon fixing , and defining and as in Lemma 12, the continuity of allows us to find such that for all . This establishes the existence of
We show . Indeed, since it holds that
for all , it follows, in conjunction with (55) and (56), that
(57) |
for all . Solving the ODI (57) yields
(58) |
with some , necessitating because if , the continuity of would imply , contradicting the growth established in (58). Hence, in accordance with (58), we obtain
which gives the desired results. In particular, in the three-dimensional case, since can be set to satisfy , it holds that
for all , which shows that must be finite, as desired. ∎
Acknowledgement
J. Ahn is supported by NRF grant No. RS-2024-00336346. K. Kang is supported by NRF grant No. RS-2024-00336346. D. Kim is supported by the National Research Foundation of Korea(NRF) grant funded by the Korea government(MSIT)(grant No. 2022R1A4A1032094).
References
- [1] J. Ahn, K. Kang, and J. Lee. Regular solutions of chemotaxis-consumption systems involving tensor-valued sensitivities and Robin type boundary conditions. Mathematical Models and Methods in Applied Sciences, 33(11):2337–2360, 2023.
- [2] J. Ahn, K. Kang, and C. Yoon. Global classical solutions for chemotaxis‐fluid systems in two dimensions. Mathematical Methods in the Applied Sciences, 44(2):2254–2264, 2021.
- [3] J. Ahn and M. Winkler. A critical exponent for blow-up in a two-dimensional chemotaxis-consumption system. Calculus of Variations and Partial Differential Equations, 62(6):180, 2023.
- [4] N. D. Alikakos. Lp bounds of solutions of reaction-diffusion equations. Communications in Partial Differential Equations, 4(8):827–868, 1979.
- [5] T. Cieślak and M. Winkler. Finite-time blow-up in a quasilinear system of chemotaxis. Nonlinearity, 21(5):1057–1076, 2008.
- [6] C. Dombrowski, L. Cisneros, S. Chatkaew, R. E. Goldstein, and J. O. Kessler. Self-concentration and large-scale coherence in bacterial dynamics. Physical Review Letters, 93(9):098103, 2004.
- [7] L. Fan and H.-Y. Jin. Global existence and asymptotic behavior to a chemotaxis system with consumption of chemoattractant in higher dimensions. Journal of Mathematical Physics, 58(1):011503, 2017.
- [8] M. Fuest, J. Lankeit, and M. Mizukami. Long-term behaviour in a parabolic–elliptic chemotaxis–consumption model. Journal of Differential Equations, 271:254–279, 2021.
- [9] J. Jiang. Eventual smoothness and exponential stabilization of global weak solutions to some chemotaxis systems. SIAM Journal on Mathematical Analysis, 51(6):4604–4644, 2019.
- [10] E. F. Keller and L. A. Segel. Initiation of slime mold aggregation viewed as an instability. Journal of Theoretical Biology, 26(3):399–415, 1970.
- [11] E. F. Keller and L. A. Segel. Model for chemotaxis. Journal of Theoretical Biology, 30(2):225–234, 1971.
- [12] D. Kim. Global solutions for chemotaxis-fluid systems with singular chemotactic sensitivity. Discrete and Continuous Dynamical Systems - B, 28(10):5380–5395, 2023.
- [13] J. Lankeit. Locally bounded global solutions to a chemotaxis consumption model with singular sensitivity and nonlinear diffusion. Journal of Differential Equations, 262(7):4052–4084, 2017.
- [14] J. Lankeit and G. Viglialoro. Global existence and boundedness of solutions to a chemotaxis-consumption model with singular sensitivity. Acta Applicandae Mathematicae, 167(1):75–97, 2020.
- [15] J. Lankeit and M. Winkler. Depleting the signal: analysis of chemotaxis-consumption models—a survey. Studies in Applied Mathematics, 151(4):1197–1229, 2023.
- [16] Y. Tao. Boundedness in a chemotaxis model with oxygen consumption by bacteria. Journal of Mathematical Analysis and Applications, 381(2):521–529, 2011.
- [17] Y. Tao and M. Winkler. Boundedness in a quasilinear parabolic-parabolic Keller-Segel system with subcritical sensitivity. Journal of Differential Equations, 252(1):692–715, 2012.
- [18] Y. Tao and M. Winkler. Eventual smoothness and stabilization of large-data solutions in a three-dimensional chemotaxis system with consumption of chemoattractant. Journal of Differential Equations, 252(3):2520–2543, 2012.
- [19] Y. Tao and M. Winkler. Global existence and boundedness in a Keller-Segel-Stokes model with arbitrary porous medium diffusion. Discrete and Continuous Dynamical Systems - A, 32(5):1901–1914, 2012.
- [20] I. Tuval, L. Cisneros, C. Dombrowski, C. W. Wolgemuth, J. O. Kessler, and R. E. Goldstein. Bacterial swimming and oxygen transport near contact lines. Proceedings of the National Academy of Sciences, 102(7):2277–2282, 2005.
- [21] G. Viglialoro. Global existence in a two-dimensional chemotaxis-consumption model with weakly singular sensitivity. Applied Mathematics Letters, 91:121–127, 2019.
- [22] L. Wang, C. Mu, K. Lin, and J. Zhao. Global existence to a higher-dimensional quasilinear chemotaxis system with consumption of chemoattractant. Zeitschrift für angewandte Mathematik und Physik, 66(4):1633–1648, 2015.
- [23] L. Wang, C. Mu, and S. Zhou. Boundedness in a parabolic-parabolic chemotaxis system with nonlinear diffusion. Zeitschrift für angewandte Mathematik und Physik, 65(6):1137–1152, 2014.
- [24] Y. Wang and M. Winkler. Finite-time blow-up in a repulsive chemotaxis-consumption system. Proceedings of the Royal Society of Edinburgh Section A: Mathematics, 153(4):1150–1166, 2023.
- [25] M. Winkler. Global large-data solutions in a chemotaxis-(Navier-)Stokes system modeling cellular swimming in fluid drops. Communications in Partial Differential Equations, 37(2):319–351, 2012.
- [26] M. Winkler. The two-dimensional Keller–Segel system with singular sensitivity and signal absorption: Global large-data solutions and their relaxation properties. Mathematical Models and Methods in Applied Sciences, 26(05):987–1024, 2016.
- [27] M. Winkler. Renormalized radial large-data solutions to the higher-dimensional Keller–Segel system with singular sensitivity and signal absorption. Journal of Differential Equations, 264(3):2310–2350, 2018.
- [28] M. Winkler. Approaching logarithmic singularities in quasilinear chemotaxis-consumption systems with signal-dependent sensitivities. Discrete and Continuous Dynamical Systems - B, 27(11):6565, 2022.
- [29] J. Yan and Y. Li. Global generalized solutions to a Keller–Segel system with nonlinear diffusion and singular sensitivity. Nonlinear Analysis, 176:288–302, 2018.
- [30] S.-O. Yang and J. Ahn. Long time asymptotics of small mass solutions for a chemotaxis-consumption system involving prescribed signal concentrations on the boundary. Nonlinear Analysis: Real World Applications, 79:104129, 2024.
- [31] Q. Zhang and Y. Li. Stabilization and convergence rate in a chemotaxis system with consumption of chemoattractant. Journal of Mathematical Physics, 56(8):081506, 2015.