Log in

The existence and stability of spikes in the one-dimensional Keller–Segel model with logistic growth

  • Published:
Journal of Mathematical Biology Aims and scope Submit manuscript

Abstract

It is well known that Keller–Segel models serve as a paradigm to describe the self aggregation phenomenon, which exists in a variety of biological processes such as wound healing, tumor growth, etc. In this paper, we study the existence of monotone decreasing spiky steady state and its linear stability property in the Keller–Segel model with logistic growth over one-dimensional bounded domain subject to homogeneous Neumann boundary conditions. Under the assumption that chemo-attractive coefficient is asymptotically large, we construct the single boundary spike and next show this non-constant steady state is locally linear stable via Lyapunov–Schmidt reduction method. As a consequence, the multi-symmetric spikes are obtained by reflection and periodic extension. In particular, we present the formal analysis to illustrate our rigorous theoretical results.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Subscribe and save

Springer+ Basic
EUR 32.99 /Month
  • Get 10 units per month
  • Download Article/Chapter or Ebook
  • 1 Unit = 1 Article or 1 Chapter
  • Cancel anytime
Subscribe now

Buy Now

Price includes VAT (Germany)

Instant access to the full article PDF.

Fig. 1
Fig. 2

Similar content being viewed by others

References

  • Carrillo J, Li J, Wang Z-A (2021) Boundary spike-layer solutions of the singular Keller–Segel system: existence and stability. Proc Lond Math Soc 122(1):42–68

    MATH  Google Scholar 

  • Chen L, Kong F, Wang Q (2020) Stationary ring and concentric-ring solutions of the Keller–Segel model with quadratic diffusion. SIAM J Math Anal 52(5):4565–4615

    MATH  Google Scholar 

  • Chen L, Kong F, Wang Q (2021) Global and exponential attractor of the repulsive Keller–Segel model with logarithmic sensitivity. Eur J Appl Math 32(4):599–617

    MATH  Google Scholar 

  • Chen X, Hao J, Wang X, Wu Y, Zhang Y (2014) Stability of spiky solution of Keller–Segel’s minimal chemotaxis model. J Differ Equ 257(9):3102–3134

    MATH  Google Scholar 

  • Childress S, Percus K (1981) Nonlinear aspects of chemotaxis. Math Biosci 56(3–4):217–237

    MATH  Google Scholar 

  • del Pino M, Wei J (2006) Collapsing steady states of the Keller–Segel system. Nonlinearity 19(3):661

    MATH  Google Scholar 

  • del Pino M, Wei J (2016) An introduction to the finite and infinite dimensional reduction method. Geom Anal Around Scalar Curvatures 31:35–118

    MATH  Google Scholar 

  • FlexPDE. PDE solutions inc. https://www.pdesolutions.com, (2021)

  • Gui C (1996) Multipeak solutions for a semilinear Neumann problem. Duke Math J 84(3):739–769

    MATH  Google Scholar 

  • Gui C, Wei J (1999) Multiple interior peak solutions for some singularly perturbed Neumann problems. J Differ Equ 158(1):1–27

    MATH  Google Scholar 

  • Gui C, Wei J, Winter M (2000) Multiple boundary peak solutions for some singularly perturbed Neumann problems. Ann Henri Poincaré 17:47–82

    MATH  Google Scholar 

  • Herrero M, Velázquez J (1996) Chemotactic collapse for the Keller–Segel model. J Math Biol 35(2):177–194

    MATH  Google Scholar 

  • Hillen T, Painter K (2009) A user’s guide to PDE models for chemotaxis. J Math Biol 58(1):183–217

    MATH  Google Scholar 

  • Hillen T, Potapov A (2004) The one-dimensional chemotaxis model: global existence and asymptotic profile. Math Methods Appl Sci 27(15):1783–1801

    MATH  Google Scholar 

  • Hillen T, Zielinski J, Painter K (2013) Merging-emerging systems can describe spatio-temporal patterning in a chemotaxis model. Discret Contin Dyn Syst Ser B 18(10):2513

    MATH  Google Scholar 

  • Horstmann D (2001) The nonsymmetric case of the Keller–Segel model in chemotaxis: some recent results. Nonlinear Differ Equ Appl 8(4):399–423

    MATH  Google Scholar 

  • Horstmann D (2003) From 1970 until present: the Keller–Segel model in chemotaxis and its consequences i. Jahresber Deutsch Math-Verein 105:103–165

    MATH  Google Scholar 

  • Horstmann D (2004) From 1970 until present: the Keller–Segel model in chemotaxis and its consequences ii. Jahresber Deutsch Math-Verein 106:51–69

    MATH  Google Scholar 

  • Horstmann D, Winkler M (2005) Boundedness versus blow-up in a chemotaxis system. J Differ Equ 215(1):52–107

    MATH  Google Scholar 

  • ** L, Wang Q, Zhang Z (2016) Pattern formation in Keller–Segel chemotaxis models with logistic growth. Int J Bifurcat Chaos 26(02):1650033

    MATH  Google Scholar 

  • Kang K, Kolokolnikov T, Ward M (2007) The stability and dynamics of a spike in the 1D Keller–Segel model. IMA J Appl Math 72(2):140–162

    MATH  Google Scholar 

  • Keller E, Segel L (1970) Initiation of slime mold aggregation viewed as an instability. J Theor Biol 26(3):399–415

    MATH  Google Scholar 

  • Keller E, Segel L (1971) Model for chemotaxis. J Theoret Biol 30(2):225–234

    MATH  Google Scholar 

  • Kolokolnikov T, Wei J, Alcolado A (2014) Basic mechanisms driving complex spike dynamics in a chemotaxis model with logistic growth. SIAM J Appl Math 74(5):1375–1396

    MATH  Google Scholar 

  • Kolokolnikov T, Wei J, Winter M (2009) Existence and stability analysis of spiky solutions for the Gierer–Meinhardt system with large reaction rates. Phys D 238(16):1695–1710

    MATH  Google Scholar 

  • Kong F, Wang Q (2022) Stability, free energy and dynamics of multi-spikes in the minimal Keller–Segel model. Discret Contin Dyn Syst Ser B 42(5):2499

    MATH  Google Scholar 

  • Kong F, Wei J, Xu L (2022) Existence of multi-spikes in the Keller–Segel model with logistic growth. preprint

  • Lin C-S, Ni W-M, Takagi I (1988) Large amplitude stationary solutions to a chemotaxis system. J Differ Equ 72(1):1–27

    MATH  Google Scholar 

  • Lin F-H, Ni W-M, Wei J (2007) On the number of interior peak solutions for a singularly perturbed Neumann problem. Commun Pure Appl Math 60(2):252–281

    MATH  Google Scholar 

  • Lin K, Mu C (2017) Convergence of global and bounded solutions of a two-species chemotaxis model with a logistic source. Discret Contin Dyn Syst Ser B 22(6):2233

    MATH  Google Scholar 

  • Nagai T (1995) Blow-up of radially symmetric solutions to a chemotaxis system. Adv Math Sci Appl 5:581–601

    MATH  Google Scholar 

  • Nagai T, Senba T, Suzuki T (2000) Chemotactic collapse in a parabolic system of mathematical biology. Hiroshima Math J 30(3):463–497

    MATH  Google Scholar 

  • Nagai T, Senba T, Yoshida K (1997) Application of the Trudinger–Moser inequah. ty to a parabolic system of chemotaxis. Funkc Ekvacioj 40:411–433

    MATH  Google Scholar 

  • Nanjundiah V (1973) Chemotaxis, signal relaying and aggregation morphology. J Theor Biol 42(1):63–105

    Google Scholar 

  • Ni W-M, Takagi I (1991) On the shape of least-energy solutions to a semilinear Neumann problem. Commun Pure Appl Math 44(7):819–851

    MATH  Google Scholar 

  • Ni W-M, Takagi I (1993) Locating the peaks of least-energy solutions to a semilinear Neumann problem. Duke Math J 70(2):247–281

    MATH  Google Scholar 

  • Novick-Cohen A, Segel L (1984) A gradually slowing travelling band of chemotactic bacteria. J Math Biol 19(1):125–132

    MATH  Google Scholar 

  • Osaki K, Yagi A (2001) Finite dimensional attractor for one-dimensional Keller–Segel equations. Funkc Ekvacioj 44(3):441–470

    MATH  Google Scholar 

  • Painter K, Hillen T (2011) Spatio-temporal chaos in a chemotaxis model. Phys D 240(4–5):363–375

    MATH  Google Scholar 

  • Schaaf R (1985) Stationary solutions of chemotaxis systems. Trans Am Math Soc 292(2):531–556

    MATH  Google Scholar 

  • Senba T, Suzuki T (2000) Some structures of the solution set for a stationary system of chemotaxis. Adv Math Sci Appl 10(1):191–224

    MATH  Google Scholar 

  • Wang G, Wei J (2002) Steady state solutions of a reaction-diffusion system modeling chemotaxis. Math Nachr 233(1):221–236

    MATH  Google Scholar 

  • Wang Q, Yan J, Gai C (2016) Qualitative analysis of stationary Keller–Segel chemotaxis models with logistic growth. Z Angew Math Phys 67(3):1–25

    MATH  Google Scholar 

  • Wang X, Xu Q (2013) Spiky and transition layer steady states of chemotaxis systems via global bifurcation and Helly’s compactness theorem. J Math Biol 66(6):1241–1266

    MATH  Google Scholar 

  • Wei J (1998) On the interior spike solutions for some singular perturbation problems. Proc R Soc Edinb A 128(4):849–874

    MATH  Google Scholar 

  • Wei J, Winter M (1998) Stationary solutions for the Cahn–Hilliard equation. Ann Henri Poincaré. 15:459–492

    MATH  Google Scholar 

  • Wei J, Winter M (2014) Mathematical aspects of pattern formation in biological systems, vol 189. Springer, London

    MATH  Google Scholar 

  • Winkler M (2008) Chemotaxis with logistic source: very weak global solutions and their boundedness properties. J Math Anal Appl 348(2):708–729

    MATH  Google Scholar 

  • Winkler M (2010) Boundedness in the higher-dimensional parabolic–parabolic chemotaxis system with logistic source. Commun Part Differ Equ 35(8):1516–1537

    MATH  Google Scholar 

  • Winkler M (2014) Global asymptotic stability of constant equilibria in a fully parabolic chemotaxis system with strong logistic dampening. J Diff Equ 257(4):1056–1077

    MATH  Google Scholar 

  • ** can prevent blow-up for the minimal Keller–Segel chemotaxis system? J Math Anal Appl 459(2):1172–1200

    MATH  Google Scholar 

  • **ang T (2018) Sub-logistic source can prevent blow-up in the 2D minimal Keller–Segel chemotaxis system. J Math Phys 59(8):081502

    MATH  Google Scholar 

  • Zhang Y, Chen X, Hao J, Lai X, Qin C (2014) An eigenvalue problem arising from spiky steady states of a minimal chemotaxis model. J Math Anal Appl 420(1):684–704

    MATH  Google Scholar 

Download references

Acknowledgements

The research of J. Wei is partially supported by NSERC of Canada. L. Xu would like to express the gratitude for the financial support of China Scholarship Council and Natural Science Foundation of China (No. 11931012).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Fanze Kong.

Ethics declarations

Conflict of interest

The authors declare that there are no conflicts of interest regarding the publication of this paper.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix A: Formal expansion of single boundary spikes

In Appendix A, we shall employ the matched asymptotic analysis to reconstruct the non-constant steady state (1.4), which can support our rigorous argument.

We firstly multiply the both hand side of the u-equation by \(\frac{1}{\chi }\) in (1.3) to obtain

$$\begin{aligned} \left\{ \begin{array}{ll} 0=\frac{1}{\chi }u_{xx}- (uv_x)_x+\frac{\mu }{\chi }u({{\bar{u}}}-u),&{}x\in (0,L),\\ 0=v_{xx}-v+u,&{}x\in (0,L). \end{array} \right. \end{aligned}$$
(1.1)

Our aim is to look for a localized pattern with the centre being 0. Recall \(\epsilon :=\sqrt{1/\chi }\), then in the inner region, we introduce

$$\begin{aligned} U(y):=u(x),\quad V(y):=v(x)\text {~with~} y:=\frac{x }{\epsilon }. \end{aligned}$$
(1.2)

Upon substituting (1.2) into (1.1), one has

$$\begin{aligned} \left\{ \begin{array}{ll} 0=U_{yy}-\epsilon ^{-2} (U V_y)_y+\epsilon ^2\mu U({{\bar{u}}}-U),&{}y\in (0,\frac{L}{\epsilon }),\\ 0=\epsilon ^{-2} V_{yy}-V+U,&{}y\in (0,\frac{L}{\epsilon }). \end{array} \right. \end{aligned}$$
(1.3)

We expand

$$\begin{aligned} U(y)=U_0+\epsilon ^{2}U_1+\ldots , \quad V(y)= V_0+\epsilon ^{2} V_1\cdots \end{aligned}$$
(1.4)

and substitute it into (1.3). To the leading order, we see that \( V_{0yy}=0\). Since we would like to find the uniformly bounded solution in \((0,\infty )\), one takes \(V_0=V_{00}\), where \(V_{00}\) is an undetermined constant. Moreover, we collect the following hierarchy from (1.3) and (1.4):

$$\begin{aligned} \left\{ \begin{array}{ll} U_{0yy}-(U_0 V_{1y})_y=0, &{}y\in (0,\infty ),\\ V_{1yy}-V_{00}+U_0=0, &{}y\in (0,\infty ), \end{array} \right. \end{aligned}$$
(1.5)

and

$$\begin{aligned} \left\{ \begin{array}{ll} U_{1yy}-(U_1V_{1y})_y-(U_0V_{2y})_y=-\mu U_0({{\bar{u}}}-U_0),&{}y\in (0,\infty ),\\ V_{2yy}=V_1-U_1,&{}y\in (0,\infty ). \end{array} \right. \end{aligned}$$

Noting \(U\ll 1\) in the outer region, one has \(U_0\rightarrow 0\) as \(\vert y\vert \rightarrow 0\). Thus, we infer from the first equation of (1.5) that \(U_0(y)=U_{00} e^{ V_1(y)}\), where \(U_{00}\) is an unknown constant. Now in the outer region, we can replace U in sense of distribution by

$$\begin{aligned} U\rightarrow \epsilon U_{00}\int _0^\infty e^{V_1}d\rho \delta _0(x). \end{aligned}$$

As such we find the outer problem for v is

$$\begin{aligned} \left\{ \begin{array}{ll} v_{yy}-v=-\epsilon U_{00}C\delta _0(x), &{}x\in (0,L),\\ v_{y}(0)=v_{y}(L)=0, \end{array} \right. \end{aligned}$$

where \(C:=\int _0^\infty e^{V_1}d\rho \) and we impose that C is a finite integral. To express v in the outer region, we introduce the following one-dimensional Neumann Green’s function \(G(x;\xi )\):

$$\begin{aligned} \left\{ \begin{array}{ll} G_{xx}-G=-\delta (x-\xi ), &{}x\in (0,L),\\ G_x(0;\xi )=G_x(L;\xi )=0, \end{array} \right. \end{aligned}$$

where G has the following explicit form:

$$\begin{aligned} G(x;\xi )=\left\{ \begin{array}{ll} \frac{\cosh (L-\xi )}{\sinh L}\cosh x,&{}x\in (0,\xi ),\\ \frac{\cosh \xi }{\sinh L}\cosh (L-x), &{}x\in (\xi ,L). \end{array} \right. \end{aligned}$$

Hence, we find v satisfies \(v\sim \epsilon U_{00}CG(x;0)\). It follows that as \(x\rightarrow 0\) and \(\epsilon \rightarrow 0^+\), \(v\rightarrow 0\). Recall in the inner expansion, \(V=V_{00}+\epsilon ^2 V_1+\cdots ,\) then we conclude from the matching that \(V_{00}=0.\)

We next solve (1.5) to find the inner solution. Upon substituting \(U_0=U_{00}e^{V_1}\) and \(V_{00}=0\) into the second equation, we establish the following core problem:

$$\begin{aligned} \left\{ \begin{array}{ll} V_{1yy}+U_{00} e^{V_1}=0, ~y\in (0,\infty ),\\ V_{1y}(0)=0. \end{array} \right. \end{aligned}$$
(1.6)

Solving equation (1.6) gives rise to

$$\begin{aligned} V_1(y)=\log \Bigg (\frac{a}{2 U_{00} \cosh ^2(\frac{\sqrt{a}}{2} y)}\Bigg ), \end{aligned}$$

where a is a free parameter. Since \( V_{1y}=-\sqrt{a}\tanh \big (\frac{\sqrt{a}}{2} y\big )\), one has from the relationship between \(V_{1}\) and \(U_{0}\) that

$$\begin{aligned} U_0 =\frac{a}{2}\textrm{sech}^{2}\Big (\frac{\sqrt{a}}{2} y\Big ). \end{aligned}$$
(1.7)

To determine the constants, we next apply the integral constraint \(\int _{0}^L u({{\bar{u}}}-u)dx=0\) thanks to the Neumann boundary condition. Noting \(U\sim U_0\), we can arrive at

$$\begin{aligned}\int ^\infty _{0}U_0({{\bar{u}}}-U_0)dy=0.\end{aligned}$$

By using \(U_0(y)=U_{00} e^{V_1}\), one has \(\int _{0}^{\infty }(e^{V_1}-U_{00} e^{2 V_1})dy=0\), which then yields that

$$\begin{aligned} \int _{0}^{\infty }\Bigg ({{\bar{u}}}\textrm{sech}^{2}\Bigg (\frac{\sqrt{a}}{2} y\Bigg ) -\frac{a}{2}\textrm{sech}^{4}\Bigg (\frac{\sqrt{a}}{2} y\Bigg )\Bigg )dy=0. \end{aligned}$$
(1.8)

Let \(z=\frac{\sqrt{a}}{2} y\), then we solve (1.8) to get \(a\sim 3{{\bar{u}}}.\) Therefore, we have from (1.7) that for \(y\in (0,\infty ),\)

$$\begin{aligned} U_0=\frac{a}{2}\textrm{sech}^{2}\Big (\frac{\sqrt{a}}{2} y\Big ) \end{aligned}$$

with \(a\sim 3{{\bar{u}}}.\) Now, we have obtained the inner solution.

Focusing on the outer region, one has \(u\sim 0\) and \(v_{xx}-v\sim 0\) with \(v_{0y}(L)=0\). By solving it, we get

$$\begin{aligned}v\sim C_v \cosh (x-L),~~x\in (0,L),\end{aligned}$$

where \(C_v\) is an unknown constant to be determined. We next use the matching condition to determine constant \(C_v\). From the inner solution, one finds

$$\begin{aligned} \frac{d V}{dy}=-\epsilon ^{2}\sqrt{a}\tanh \Bigg (\frac{\sqrt{a}}{2}y\Bigg ), \end{aligned}$$

which yields

$$\begin{aligned} \frac{d V}{dy}\rightarrow -\epsilon ^{2}\sqrt{a}\text {~~as~~} y\rightarrow +\infty . \end{aligned}$$
(1.9)

On the other hand, from the outer solution, we conclude for \(0<x<L,\)

$$\begin{aligned}\frac{dv}{dy}=\frac{dv}{dx}\epsilon \sim \epsilon C_v \sinh (x-L).\end{aligned}$$

It follows that

$$\begin{aligned} \frac{dv}{dy}\rightarrow -C_v \epsilon \sinh L~~\text {as}~~x\rightarrow 0. \end{aligned}$$
(1.10)

After matching (1.9) and (1.10), one can get \(C_{\nu }=\epsilon \sqrt{a}/\sinh L. \)

In summary, the single boundary spike \((u^-,v^-)\) can be asymptotically written as

$$\begin{aligned} u^{-}\sim \frac{a}{2}\textrm{sech}^2\bigg (\frac{\sqrt{a}}{2}\cdot \frac{x}{\epsilon }\Bigg ),~\text {~for}~~x\in (0,L), \end{aligned}$$

and

$$\begin{aligned} v^{-}\sim \epsilon \frac{\sqrt{a}}{\sinh L} \cosh (x-L),\text {~~for~~}x\in (0,L), \end{aligned}$$

where \(a\sim 3{{\bar{u}}}\). Next, we use the Van Dyke’s matching principle \(v_\text {unif}=v_\text {inner}+v_\text {outer}-v_\text {overlap}\) to find the composite expansion of v, which is

$$\begin{aligned}v^-\sim \epsilon ^2\log \Big (\frac{1}{4}\textrm{sech}^2\Big (\frac{\sqrt{a}}{2}\frac{x}{\epsilon }\Big )\Big )+\epsilon \frac{\sqrt{a}}{\sinh L} \cosh (x-L)+\sqrt{a}\epsilon x.\end{aligned}$$

These results agree with those stated in Theorem 1.1.

Appendix B: Formal analysis of the eigenvalue problem

This section is devoted to the study of linearized eigenvalue problem (4.1) via the matched asymptotic analysis. Similarly, in the inner region, we introduce the following rescaled functions:

$$\begin{aligned}\Phi (y):=\phi (x),~~\Psi (y):=\psi (x)\text {~with~} y=\frac{x}{\epsilon }.\end{aligned}$$

By using it together with (1.2), one can rewrite (4.1) as

$$\begin{aligned} \left\{ \begin{array}{ll} \epsilon ^2\lambda \Phi =\Phi _{yy}-\epsilon ^{-2} (U \Psi _y+\Phi V_{y})_y+\epsilon ^2\mu ({{\bar{u}}}-2U)\Phi ,&{}y\in (0,\frac{L}{\epsilon }),\\ \lambda \Psi =\epsilon ^{-2}\Psi _{yy}-\Psi +\Phi , &{}y\in (0,\frac{L}{\epsilon }),\\ \Phi _y(0)=\Psi _y(0)=\Phi _y(\frac{L}{\epsilon })=\Psi _y(\frac{L}{\epsilon })=0. \end{array} \right. \end{aligned}$$
(2.1)

Similarly as above, we expand

$$\begin{aligned} \lambda =\lambda _0+\cdots ,~~\Phi (y)= \Phi _0(y)+\epsilon ^{2}\Phi _1(y)+\cdots \text {~~and~~} \Psi = \Psi _0(y)+\epsilon ^{2}\Psi _1(y)+\cdots \end{aligned}$$

and substitute them together with (1.4) into (2.1). Then one can find \(\Psi _{0y}=0\), and thereby \(\Psi _0(y):=\Psi _{00}\) with \(\Psi _{00}\) being a constant. Moreover, with the help of matching condition between the inner and the outer solution, we obtain \(\Psi _{00}=0.\)

We further collect the following leading order system:

$$\begin{aligned} \left\{ \begin{array}{ll} 0=\Phi _{0yy}-(U_0\Psi _{1y}+\Phi _0 V_{1y})_y,&{}y\in (0,\infty ),\\ 0=\Psi _{1yy}+\Phi _0,&{}y\in (0,\infty ),\\ \Phi _{0y}(0)=\Psi _{1y}(0)=0,~~\Phi _0(\infty )=0. \end{array} \right. \end{aligned}$$
(2.2)

The first equation in (2.2) implies that \(\Big (\frac{\Phi _0}{U_0}\Big )_y=\Psi _{1y}\), hence \(\Phi _0=U_0\Psi _1+CU_0\) thanks to the boundary conditions, where C is some constant to be determined later on. Therefore, (2.2) yields that

$$\begin{aligned} \left\{ \begin{array}{ll} \Psi _{1yy}+U_0\Psi _1+CU_0=0,&{}y\in (0,\infty ),\\ \Psi _{1y}(0)=0. \end{array} \right. \end{aligned}$$
(2.3)

Since \(U_0= \frac{a}{2}\textrm{sech}^{2}\Big (\frac{\sqrt{a}}{2}y\Big )\) and \(a\sim 3{{\bar{u}}}\), we further solve (2.3) to get the eigenfunctions are unique up to a constant multiplier of the following

$$\begin{aligned} \Psi _1(y)=&\,\frac{1}{\sqrt{a}}y\tanh \Big (\frac{\sqrt{a}}{2} y\Big ),\nonumber \\ \Phi _0(y)=&\Bigg (1-\frac{\sqrt{a}}{2}y\tanh \Bigg (\frac{\sqrt{a}}{2}y\Bigg )\Bigg )\textrm{sech}^{2}\Bigg (\frac{\sqrt{a}}{2}y\Bigg ),~~a\sim 3{{\bar{u}}}. \end{aligned}$$
(2.4)

Next, we proceed to show the corresponding leading eigenvalue \(\lambda _0<0\), which tells us that steady state (1.4) is linearly stable.

Proof

We integrate the \(\phi \)-equation in (4.1) over (0, L) to get

$$\begin{aligned} \lambda \int _{0}^L\phi (x) dx=\mu \int _0^L ({{\bar{u}}}-2u)\phi (x) dx. \end{aligned}$$
(2.5)

Upon substituting (2.4) into (2.5), one has the left hand side and right hand side satisfy

$$\begin{aligned} \lambda \int _0^L\phi (x) dx\sim \epsilon \lambda _0 \int _{0}^\infty \Phi _0(y) dy \end{aligned}$$
(2.6)

and

$$\begin{aligned} \int _0^L \mu ({{\bar{u}}}-2u)\phi (x) dx\sim \epsilon \int _{0}^\infty \mu ({{\bar{u}}}-2 U_0) \Phi _0(y) dy, \end{aligned}$$
(2.7)

respectively. By straightforward calculation, we obtain

$$\begin{aligned} \int _0^\infty \Phi _0 dy=-\frac{1}{\sqrt{a}}, \end{aligned}$$
(2.8)

and

$$\begin{aligned} \int _0^\infty U_0 \Phi _0 dy=-\frac{1}{2}\sqrt{a}. \end{aligned}$$
(2.9)

Combining (2.8) and (2.9), we have from (2.5), (2.6) and (2.1) that \(\lambda _0\sim -2\mu {{\bar{u}}}.\) This gives us (1.4) is linearly stable with respect to the even eigenfunction (2.4), then formally verifies Theorem 1.2. \(\square \)

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Kong, F., Wei, J. & Xu, L. The existence and stability of spikes in the one-dimensional Keller–Segel model with logistic growth. J. Math. Biol. 86, 6 (2023). https://doi.org/10.1007/s00285-022-01840-1

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s00285-022-01840-1

Keywords

Mathematics Subject Classification

Navigation