1 Main results and discussion

In this work, the main object of study is the equation:

$$\begin{aligned} u_{t}(t,x)= & {} u_{xx}(t,x) - u(t,x) + g(u(t-h,x)),\quad x \in {\mathbb {R}},\ t >0, \end{aligned}$$
(1)

where \(h>0\) and the nonlinear reaction term \(g:{\mathbb {R}}_+\rightarrow {\mathbb {R}}_+\) is Lipschitz continuous with Lipschitz constant \(L_g\) which satisfies the monostability condition \(L_g\ge g'(0)>1\) and has exactly two fixed points: 0 and \(\kappa >0.\) The Eq. (1) is frequently considered to model problems of population dynamics. In this case g stands for the birth rate function, h is the age when the individual reaches the sexual maturity and u(tx) is the adult population at location x and time t. The diffusion and death rates have been normalized. In this framework is relevant the stability properties of the positive equilibrium \(\kappa \) and the existence of colonization waves so-called wavefronts (see [2, 12, 21, 28, 34, 35] and references therein). Wavefronts with speed c are non-negative entire bounded solutions \(u(t,x)=\psi _c(x+ct)\) such that the profile \(\psi _c:{\mathbb {R}}\rightarrow {\mathbb {R}}\) satisfies \(\psi _c(-\infty )=0\) and \(\psi _c(+\infty )=\kappa .\) It is well known that when g is monotone and \(h\ge 0\) then there exists a positive number \(c_*=c_*(h)\) so-called critical speed or minimal speed such that (1) has wavefronts if and only if \(c\ge c_*\) [18, 40, 43]; wavefronts with speed \(c_*\) are called critical wavefronts. Moreover, these wavefronts are monotone and unique modulo translation. The main tool to obtain the stability and existence of wavefronts is to construct sub and super- solutions by using monotony arguments.

However, when g is non-monotone the associated semi-flow is non-monotone in general and wavefronts are replaced by positive bounded solutions \(u(t, x)=\psi _c(x+ct)\) such that \(\psi _c(-\infty )=0\) and \(\liminf _{x\rightarrow +\infty }\psi _c(x)>0\) which are called semi-wavefronts. For g satisfying the subtangency condition \(g(u)\le g'(0)u\), for all \(u\ge 0\), has been demonstrated the existence of a minimal speed \(c_*=c_*(h)\) for the existence of semi-wavefronts to (1) for all \(h\ge 0\) (see [37, Theorems 4.5 and 5.4] and [13, 43]). Under the Diekmann- Kaper (D-K, for short) condition \(L_g=g'(0)\) (see [11, Theorem 6.4]) Aguerrea, Gomez and Trofimchuk demonstrated the uniqueness modulo translation of all semi-wavefronts of (1). In the general case \(L_g\ge g'(0)\) it is necessary to consider the following characteristic equation

$$\begin{aligned} E_c(\lambda ):=\lambda ^2-c\lambda -1+L_g\, e^{-\lambda ch}=0, \end{aligned}$$
(2)

for which it has been showed that there exits a speed \({\mathfrak {c}}_*={\mathfrak {c}}_*(L_g)\) defined as

$$\begin{aligned} {\mathfrak {c}}_{*}={\mathfrak {c}}_*(L_g):=\inf \{c>0: E_c(\lambda ) \, \hbox {has a positive root} \}, \end{aligned}$$
(3)

such that \(E_c\) has exactly two positive zeros \(\lambda _1(c)\le \lambda _2(c)\), also \(\lambda _1(c)=\lambda _2(c)\) if and only if \(c={\mathfrak {c}}_{*}\) (for a more detailed study of (2) see [16, Lemma 22]). Thus, the authors in [1, Theorem 4] showed that for \(c\ge {\mathfrak {c}}_{*}\) semi-wavefronts have the following representation

$$\begin{aligned} \psi _c(z)= A_{\psi _c}(-z)^{j_c}e^{\lambda _1(c) z} + e^{(\lambda _1(c)+\epsilon )z}\, r(z), \end{aligned}$$
(4)

where \(A_{\psi _c}, \epsilon \in {\mathbb {R}}_+\), \(r\in L^2({\mathbb {R}})\cap C({\mathbb {R}})\) and \(j_c= 0, 1\) with \(j_c=1\) if and only if \(L_g=g'(0)\), moreover, semi-wavefronts are unique (modulo translation) for all \(c> {\mathfrak {c}}_{*}\) and \(h\ge 0\) [1, Theorem 8]. We should mention that when \(L_g\) in (2) is replaced by \(g'(0)\) the speed \({\mathfrak {c}}_{*}\) in (3) coincides with the definition of the so-called linear speed \(c_{\#}\) and

$$\begin{aligned} c_{\#}\le c_{*}\le {\mathfrak {c}}_{*}, \end{aligned}$$
(5)

(see [37, Theorem 4.5 and Theorem 5.4]), also when g is subtangential then \(c_{\#}\) coincides with the critical speed \(c_*\) for the existence of semi-wavefronts (see [43, Theorem 4.4]). In particular, if g satisfies the D-K condition \(L_g=g'(0)\) then g is subtangential and \({\mathfrak {c}}_{*}=c_*=c_{\#}\). One of the main results of this work is show the stability of semi-wavefronts (with unbounded exponential weights) for \(c\ge {\mathfrak {c}}_{*}(h)\) for all delay \(h\ge 0\) (the same conditions for c and h to establish the uniqueness of semi-wavefronts in [1]).

In order to overcome the non-monotony of some reaction–diffusion equations with delay a quasi-monotonicity condition is assumed which usually requires the monotony of the delayed argument. Indeed, in a pioneering work, Schaaf [29] considered the following parabolic functional differential equation

$$\begin{aligned} u_t(t, x)=u_{xx}(t, x)+f(u(t, x), u(t-h, x)), \quad x\in {\mathbb {R}}, \, t>0. \end{aligned}$$
(6)

Schaaf proved that for a concave nonlinearity f with exactly two equilibria (0 and 1) satisfying a certain positivity condition (see [29, Section 2.1]) and

$$\begin{aligned} \partial _2f(u,v)\ge 0\quad \hbox {for all}\, u,v\in {\mathbb {R}}_+\quad \quad \hbox { (quasi-monotonicity condition)} \end{aligned}$$
(7)

wavefronts are linearly stable for small delay h [29, Theorem 4.13].

After the pioneering work by Schaaf, a series of other studies appeared where the KPP condition \(|g'(u)|\le g'(0),\) for all \(u\ge 0\), or the concave condition \(g''(u)\le 0\), for all \(u\ge 0\), was instrumental for the stability analysis. Among these studies, we would like to distinguish an important contribution [24] by Mei, Ou and Zhao where the authors proved the global stability of monotone wavefronts (critical as well as non-critical ones, see [24, Theorem 2.2]) of the following non-local equation

$$\begin{aligned} u_t(t,x)= u_{xx}(t,x)-u(t,x)+\int _{{\mathbb {R}}}K(y)g(u(t-h,x-y))dy\quad x\in {\mathbb {R}}, \ t>0,\nonumber \\ \end{aligned}$$
(8)

for monotone and concave g and K a heat kernel; here the perturbations are taken in weighted Sobolev spaces. At the same time, Lv and Wang [20] proved the global stability of non-critical wavefronts of (6) for monostable f (with exactly two equilibria: 0 and \(\kappa >0\)) satisfying (7) and the concavity condition: \(\partial _{ij}f(u, v)\le 0\) (\(\hbox {i, j}=1, 2\)), for all \(u, v\in [0, \kappa ]\). The authors in [20] also study (6) with non-local reaction term (which includes (8), for monotone g) and demonstrated the stability of non-critical monotone wavefronts in Sobolev spaces with exponential weights; this result can also be obtained by our approach to (1) even to non-monotone wavefronts, see Remark 17 (for the non-local equation (8) see our recent work [31]).

With respect to non-monotone wavefronts, we should mention a work of Wu et al. [41, Theorem 2.4] where the authors take some type non-monotone \(g\in C^2([0, \kappa ], {\mathbb {R}})\) (‘crossing monostable’ nonlinearity) satisfying \(|g'(\kappa )|< 1\) and prove the local stability of wavefronts with speed c for \(c>2\sqrt{2(L_g-1)}\) and for all \(h\ge 0\). Additionally, by assuming the KPP condition and \(|g'(\kappa )|\) sufficiently small they prove the local stability of wavefronts with speed c for all \(c> c_*\) and \(h\ge 0\) [41, Theorem 2.6]; here the existence of non-monotone wavefronts can be deduced, e.g., from [36] and [15]. Our second result for non-linear stability of wavefronts generalizes these results (see Remark 10). In this regard, for unimodal g (i.e., g has exactly one critical point which is the absolute maximum point) satisfying the KPP condition and \(|g'(\kappa )|<1\), Lin et al. [19] proved the local stability of non-critical wavefronts for all \(h\ge 0\) (monotone or non-monotone) which includes well-known models such as Nicholson’s model (see [15, 19, 33] and references therein) described by

$$\begin{aligned} u_t(t,x)=u_{xx}(t,x)-\delta u(t,x)+\rho u(t-h,x)e^{-u(t-h,x)},\quad t>0, x\in {\mathbb {R}}, \end{aligned}$$
(9)

where \(\rho ,\delta >0\), or the Mackey–Glass model [1, 3, 19, 23, 24] given by

$$\begin{aligned} u_t(t,x)=u_{xx}(t,x)-\tau \,u(t,x)+\frac{ab^n u(t-h,x)}{b^n+u^n(t-h,x)},\quad t>0, x\in {\mathbb {R}}. \end{aligned}$$
(10)

where \(\tau ,a,b >0\) and \(n\in {\mathbb {Z}}_+\). These stability results were stablished in weighted Sobolev spaces to initial data with a suitable convergence to \(\kappa \) at \(x=+\infty \).

When \(\rho /\delta \in (1, e]\) in (9) wavefronts are monotone and by [24] they (critical as well as non-critical ones) are globally stable. The authors in [19] proved the local stability of (monotone and non-monotone) non-critical wavefronts to (9) when \(\rho /\delta \in (e, e^2)\) for all \(h\ge 0\) and for small delay h when \(\rho /\delta \in (1, +\infty )\). Then, assuming \(|g'(\kappa )|<1\), Chern et al. [8, Theorem 2.3] have demonstrated the local stability of critical wavefronts (monotone or non-monotone) in the same Sobolev spaces.

Next, for Lipschitz continuous function g satisfying the D-K condition \(L_g= g'(0)\) and \(|g'(u)|<1\) in some neighborhood of \(\kappa \), Solar and Trofimchuk have established the global stability of (monotone or non-monotone) non-critical wavefronts [33, Corollary 3]. In particular, they obtained the global stability of non-critical wavefronts for (9) when \(\rho /\delta \in (1,e^2)\) for all \(h\ge 0\). Here initial data are not required to convergence to \(\kappa \) at \(x=+\infty \) as above mentioned works. Then, in a recent work, for unimodal \(g\in C^2[0, +\infty )\) satisfying the KPP condition, Mei et al. [22] have generalized the results in [8, 19] for a global perturbation in the same Sobolev spaces.

On the other hand, non-subtangential models have recently attracted a lot of interest because of their connection to the so-called Allee effect in population dynamics [6, 7, 10, 26]. More precisely, if we only consider as benefit to species a greater availability of resources then the per capita growth rate g(u) / u attains its maximum at \(u=0\), however if animal behavior is cooperative then individuals obtain benefits for intermediate densities \(u>0\) (individual fitness) which are not generated for low densities (\(u=0\)), so that the per capita growth rate g(u) / u attains its maximum at some \(u_0>0\). In this case model is said to have an Allee effect [9, Chapter 1] (since in our case the per capita growth birth rate g(u) / u is non-decreasing in a neighborhood of \(u=0\) model is said to have a weak Allee effect). In contrast to subtangential case, for a model with Allee effect it could occur \(c_*>c_{\#}\), critical wavefronts with speed \(c_*>c_{\#}\) are called pushed wavefronts. In this direction, for monotone g (necessarily non-subtangential), it has been possible to establish the stability of pushed wavefronts (see [32] and [42]) as well as that of non-critical wavefronts [33, Theorem 1]. These results show that pushed wavefronts are more attractive than critical wavefronts with speed \(c_*=c_\#\), for instance pushed wavefronts attract (orbitally) to the solution of (1) generated by the Heaviside step function while a critical wavefront (which is not a pushed wavefront) requires a logarithmic correction to attract this solution (see e.g. [12, 39] for \(h=0\) and [4] for \(h>0\)). It is important to mention that the problem of the existence of semi-wavefronts for non-subtangential models is not completely solved (e.g., see [36, Corollary 4]: of course, in the available literature there are some partial results on the existence of semi-wavefronts for certain subclasses of equations, e.g., see [37, Theorem 2.4]).

Hence, in the above mentioned works, we can find stability results for Eq. (1) only when g either is monotone or meets the sub-tangency condition. In this work we study the stability of semi-wavefronts without assuming the quasi-monotonicity nor the sub-tangency condition on g. Our approach uses ideas from [33] and a suitable Fourier analysis for partial functional differential equations. In the particular case when g is unimodal and satisfies the KPP condition, i.e. \(L_g=g'(0)\) and \(g\in C^1[0, +\infty )\), our estimates (in different spaces) are similar than [8, 19, 22, 24] for perturbations of wavefronts, but our Fourier analysis (for wavefronts and proper semi-wavefronts) is different in many aspects, for instance by our approach Fourier transforms are estimated by means of a Halanay inequality on Banach spaces (see Lemma 12) instead of finite-dimensional spaces while the non-critical case \(c>c_*\) (for wavefronts and proper semi-wavefronts) does not require Fourier analysis as in [19] (see Corollary 20 and Remark 17 below). However, approach used in [8, 19, 22, 24] allows us to obtain stability results of wavefronts on the real line when \(|g'(\kappa )|>1\) for small h whenever the initial datum \(u_0(s, x)\) converges to \(\kappa \) at \(x=+\infty \). In this regard, we obtain a general stability result for semi-wavefronts on each semi-infinite interval \((-\infty ,N]\), \(N\in {\mathbb {R}}\), without assume the restriction \(|g'(\kappa )|<1\) for all \(h\ge 0\) (Theorem 3) which also includes critical semi-wavefronts and asymptotically periodic semi-wavefronts (see, e.g [36, Theorem 3]). This kind of stability seems to be transversal to another models, indeed in a recent work [5] Benguria and Solar have stablished the stability of a class of non-monotone semi-wavefronts for the Hutchinson diffusive equation

$$\begin{aligned} u_t(t, x)= u_{xx}(t, x)+u(t,x)(1-u(t-h, x)), \quad \quad t>0, \, x\in {\mathbb {R}}, \end{aligned}$$
(11)

on each semi-infinite interval \((-\infty , N]\).

More precisely, by a suitable Fourier analysis we can show that for \(c\ge c_{*}\), an initial perturbation

$$\begin{aligned} u_0(s, x+cs)-\psi _c(x+cs)= e^{\lambda (x+cs)}\, r(s, x),\quad \hbox {for all}\quad (s, x)\in [-h, 0]\times {\mathbb {R}}, \end{aligned}$$

with \(r\in C([-h,0], L^1({\mathbb {R}}))\) and \(\lambda \) satisfying \(E_c(\lambda )\le 0\) [according to definition (2)], evolves as

$$\begin{aligned} u(t, x+ct)-\psi _c(x+ct)= O(t^{-1/2}e^{\gamma t}),\quad \hbox {for all}\quad t\ge -h, \end{aligned}$$

uniformly for \(x+ct\in (-\infty , N]\), \(N\in {\mathbb {R}}\) and some \(\gamma =\gamma (\lambda )\le 0\). Moreover, \(\gamma =0\) if and only if \(E(\lambda )=0\), i.e. \(\lambda =\lambda _1(c)\) or \(\lambda =\lambda _2(c)\).

On the other hand, in our second main result we study the convergence of perturbations of wavefronts on the remaining domain \((N, +\infty )\), \(N\in {\mathbb {R}}\). In this case it is necessary to assume the stability condition \(|g'(\kappa )|<1\) in order to establish (without assuming monotonicity or sub-tangency condition on g) the local stability of wavefronts with \(c> {\mathfrak {c}}_*\) on whole the real line \((-\infty , +\infty )\) (Corollary 20). Additionally, assuming \(|g'(u)|<1\) for u in a suitable neighborhood of \(\kappa \) we obtain the global stability of wavefronts with \(c> {\mathfrak {c}}_*\) on whole real line \((-\infty , +\infty )\) (see Theorem 9 below).

In order to obtain these stability results we study the decay of solutions of the constant coefficient linear equation with delay,

$$\begin{aligned} u_t(t,x)=u_{xx}(t,x)+mu_x(t,x)+pu(t,x)+qu(t-h,x+d), \quad x\in {\mathbb {R}}, t>0,\nonumber \\ \end{aligned}$$
(12)

where the parameters mpq y d are real numbers.

For an initial datum \(u_0\in C([-h,0], L^{1}({\mathbb {R}}))\), let us denote \(C_{u_0}:=\sup _{s\in [-h,0]}||u_0(s,\cdot )||_{L^1}\).

Theorem 1

Suppose that \(-p\ge q\ge 0\) and \(m, d\in {\mathbb {R}}\). Let \(\gamma \le 0\) be the only real solution of the following equation:

$$\begin{aligned} \gamma -p= qe^{-h\gamma }. \end{aligned}$$
(13)

If the initial datum \(u_0\) belongs to \(C ([-h,0]; L^1({\mathbb {R}}))\) then the solution u(tx) of (12) satisfies the estimate:

$$\begin{aligned} \sup _{x\in {\mathbb {R}}}|u(t,x)| <A_0 \frac{e^{\gamma t}}{\sqrt{t}}, \quad \quad \hbox {for all} \quad t>h, \end{aligned}$$
(14)

where \(A_0=C_{u_0}/2\sqrt{1+h(\gamma -p)}.\)

We note that in the special case \(-p = q\) (which implies \(\gamma = 0\)) an exponential estimate is no longer available. In some cases, it can be established that the decay is not faster than that given by (14). For instance, if \(d = 0\) for the evolution equation (12), the behavior of the solutions in the \(L^1({\mathbb {R}})\) phase space with an appropriate weight can be specified. In fact, we obtain the exact behavior which is embodied in Theorem below

Theorem 2

(Asymptotic behavior) Let us consider (12) with \(m, p\in {\mathbb {R}}\), \(q\ge 0\) and \(d = 0\). Let u(tx) be the solution generated by the initial data \(u(s,\cdot )=e^{\sigma s}u_0\) where \(u_0\) is such that \(e^{\frac{m}{2}\cdot }\,u_0\in L^1({\mathbb {R}})\) and \(\sigma \) is the only real solution of

$$\begin{aligned} qe^{-\sigma h}=\sigma +\frac{m^2}{4}-p, \end{aligned}$$
(15)

then

$$\begin{aligned} \lim _{t\rightarrow \infty }\sqrt{t}e^{-\sigma t}u(t,x+o(\sqrt{t}))=\frac{\sqrt{1+hqe^{-\sigma h}}}{2\sqrt{\pi }}e^{-\frac{m}{2}x}\int _{{\mathbb {R}}}e^{\frac{m}{2}y}u_0(y)dy, \end{aligned}$$
(16)

for all \(x\in {\mathbb {R}}\).

Now, for the study of the stability of semi-wavefronts with speed c , the following equation should be considered

$$\begin{aligned} v_t(t,z)=v_{zz}(t,z)-cv_z(t,z)-v(t,z)+g(v(t-h,z-ch)), \quad t>0, z\in {\mathbb {R}}.\nonumber \\ \end{aligned}$$
(17)

For \(c\ge {\mathfrak {c}}_*\) let us fix \(\lambda _c\in [\lambda _1(c),\lambda _2(c)]\) and let us denote by \(\xi _c(z):=e^{-\lambda _c z}\). Now, the first main result of this article can be set out.

Theorem 3

(Stability with weight) Assume that \(c\ge {\mathfrak {c}}_*\). Let \(v_0(s, z)\) and \(\psi _0 (s, z)\) be two initial data to (17) such that \(v_0,\psi _0\in C([-h,0]; L^{\infty }({\mathbb {R}})\cap C^{0,\alpha }({\mathbb {R}}))\), some \(\alpha \in (0,1]\), and

$$\begin{aligned} u_0(s,z):=\xi _c (z)|v_0(s,z)-\psi _0(s,z)|\in C([-h,0], L^{1}({\mathbb {R}})), \end{aligned}$$
(18)

then there are unique solutions v(tz) and \(\psi (t,z)\) of (17) with initial data \(v_0\) and \(\psi _0\), respectively, and these solutions satisfies \(v(\cdot +kh, \cdot ), \psi (\cdot +kh, \cdot )\in C([-h,0]; L^{\infty }({\mathbb {R}})\cap C^{0, \alpha }({\mathbb {R}}))\) for all \(k\in {\mathbb {Z}}_+\). Moreover, if u(tz) satisfies (12) with initial data \(u_0(s, z)\) and with parameters \(m=m(\lambda _c)=2\lambda _c -c\), \(p=p(\lambda _c)=\lambda _c^{2}-c\lambda _c-1\), \(q=q(\lambda _c)=L_ge^{-\lambda _c ch}\) and \(d=-ch\), then

$$\begin{aligned} \xi _c(z)|v(t,z)-\psi (t,z)|\le u(t,z),\quad \hbox {for all}\quad t\ge -h, z\in {\mathbb {R}}, \end{aligned}$$
(19)

in particular

$$\begin{aligned} |v(t,z)-\psi (t,z)|\le A_0\, \xi _c(-z)\,\frac{e^{\gamma t}}{\sqrt{t}},\quad \hbox {for all}\quad t> h, z\in {\mathbb {R}}\end{aligned}$$
(20)

where \(\gamma =\gamma (\lambda _c)\) is defined by (13) with \(p=p(\lambda _c)\) and \(q=q(\lambda _c)\).

Corollary 4

(Uniqueness) If \(\psi _c(z)\) and \(\phi _c(z)\) are two semi-wavefronts with speed \(c\ge {\mathfrak {c}}_*\) satisfying (18) then there exists \(z_*\in {\mathbb {R}}\) such that \(\psi _c(z+z_*)=\phi _c(z)\) for all \(z\in {\mathbb {R}}\).

Remark 5

If \(h=0\) in (1) then semi-wavefronts are monotone wavefronts and by taking a wavefront \(\psi (t, z)=\psi _c(z)\) in Theorem 3 we get the stability of the wavefront on the sets \((-\infty , N ], N \in {\mathbb {R}},\) which is comparable to a result obtained by Uchiyama [39, Theorem 4.1].

It has recently been showed that the estimation \(u(t, x)= O(t^{-1/2})\) in (19)—(20) for critical semi-wavefronts, in the D-K case, is actually \(u(t, x)= o(t^{-1/2})\) for all \(h\ge 0\) (see [4, Corollary 1.2]). Also, since Theorem 3 does not assume some stability condition on \(\kappa \) then semi-wavefronts could be asymptotically periodic at \(+\infty \) [36, Theorem 3] and oscillations around \(\kappa \) can be approximated by the solution v(tz) on each interval in the form \((-\infty , N]\) with \(N\in {\mathbb {R}}\). The Corollary 4 refers essentially to the fact that semi-wavefronts are equal (up to translation) if they have the same one-order asymptotic terms at \(z=-\infty \), i.e., the condition (18).

By the change of variable \(t':=\delta ^{-1} t\) and \(x':= \delta ^{-1/2}x\) Eq.  (9) can be reduced to (1) with delay \(h':=h\delta \) so that by Theorem 3 we obtain the stability of semi-wavefronts with speed c for the Nicholson’s model,

Corollary 6

(Nicholson Model) Let \(\rho /\delta \in [1,+\infty )\) be in (9). Consider \(N\in {\mathbb {R}}\) and the initial datum satisfying the conditions of Theorem 3, if \(c\ge c_*\) and \(\lambda _c\in [\lambda _1(c), \lambda _2(c)]\) then

$$\begin{aligned} \sup _{z\in (-\infty , N]}|v(t,z)-\psi _c(z)|=O(t^{-1/2}e^{\gamma t}), \end{aligned}$$

where \(\gamma =\gamma (\lambda _c)\le 0\) is determined by (13).

It is well known that if \(\rho /\delta \in (e, e^{2})\) then there are non-monotone wavefronts (see [15, Theorem 2.3]). Moreover, for some critical value \(\nu _0=2.808\ldots \) and some delay \(h_0\) if \(\rho /\delta \in [\nu _0, +\infty )\) then each minimal wavefront has oscillations around \(\kappa \) at \(+\infty \) and there exist a critical value \(c^*\) (a extended real number) such that each semi-wavefront with speed \(c>c_*\) has non-decaying slow oscillations [36, Theorem 3].

Nevertheless, the semi-wavefronts of Theorem 3 could exhibit a type of convective instability due to the positive equilibrium (e.g., see [27]), however by controlling the size of the slope of g at the positive equilibrium, the stability of the semi-wavefront on the remaining domain \([N, +\infty )\), \(N\in {\mathbb {R}}\), can be obtained. In this framework, it is necessary to assume some additional hypotheses in order to establish the existence of semi-wavefronts, such as the following condition.

(M) The function \(g:{\mathbb {R}}_+\rightarrow {\mathbb {R}}_+ \) is such that the equation \(g(x)= x\) has exactly two solutions on \([0,+\infty )\) : 0 and \(\kappa >0\). Moreover, g is \(C^1\)-smooth in some \(\delta _0\)-neighborhood of the equilibria where \(g'(0)>1>g'(\kappa ).\) In addition, there are \(C >0,\ \theta \in (0,1],\) such that \( \left| g'(u)- g'(0)\right| +|g'(\kappa ) - g'(\kappa -u)| \le Cu^\theta \) for \(u\in (0,\delta _0].\)

We note that for g satisfying (M), there are real numbers \(0 < \zeta _1\le \zeta _2\) such that

  1. (B1)

    \(g([\zeta _1,\zeta _2])\subset [\zeta _1,\zeta _2]\) and \(g([0,\zeta _1])\subset [0,\zeta _2]\);

  2. (B2)

    \(\min _{\zeta \in [\zeta _1,\zeta _2]}g(\zeta )=g(\zeta _1)\);

  3. (B3)

    \(g(x)>x\) for \(x\in [0,\zeta _1]\) and \(1<g'(0)\le g^*_+:=\sup _{s\ge 0}g(s)/s<\infty \);

  4. (B4)

    In \([0,\zeta _2]\), the equation \(g(x)=x\) has exactly two solutions 0 and \(\kappa \).

Thus, from [37, Theorem 4.5] we obtain the following result to the existence of semi-wavefronts.

Proposition 7

(Existence of semi-wavefronts) Let g satisfy (M). Then, for each \(c > {\mathfrak {c}}_*(g^*_+)\) (according to definition (3)) Eq, (1) has semi-wavefronts with speed c. Moreover, if \(\, 0 < \zeta _1\le \zeta _2\) meet (B1)–(B4) then each semi-wavefront \(\psi _c\) satisfies:

$$\begin{aligned} \zeta _1\le \liminf _{z\rightarrow -\infty }\psi _c(z)\le \limsup _{z\rightarrow +\infty }\psi _c(z)\le \zeta _2,\quad \hbox {for all} \ z\in {\mathbb {R}}. \end{aligned}$$

Remark 8

(Minimal speed for semi-wavefronts) Due [16, Theorem 18], in the case that \(g^*_+=g'(0)\) the number \({\mathfrak {c}}_*(g^*_+)\) is actually the minimal speed for the existence of semi-wavefronts.

Now, let us introduce some notation. If \(I\subset {\mathbb {R}}_+ = Dom(g)\), let us denote by

$$\begin{aligned} L_g(I):=\sup _{x\ne y;\, x,y\in I}\frac{|g(x)-g(y)|}{|x-y|}, \end{aligned}$$

and for \(b \in {\mathbb {R}}\), let us denote by \(\eta _b(z)=\min \{1,e^{\lambda _c(z-b)}\}\), with \(\lambda _c\in [\lambda _1(c), \lambda _2(c)]\). With these notations, the second main result of this paper can be established

Theorem 9

(Global stability) Let \(c> {\mathfrak {c}}_*\) and \({\bar{g}}\) be a non decreasing function satisfying (M) with equilibrium K such that \({\bar{g}}(u)\ge g(u)\) for all \(u\in {\mathbb {R}}_+\) such that \(L_{{\bar{g}}}\le L_g\). We denote by \(m_K=\min _{u\in [\kappa ,K]}g(u)\) and \({\mathcal {I}}_K:=[m_K,K]\) and we suppose that \(L_g({\mathcal {I}}_K)<1\). If for some \(q_0>0\) and \(z_0\in {\mathbb {R}}\) the initial datum satisfies

$$\begin{aligned} v_0(s,z)\ge q_0\quad \hbox {for all}\quad (s,z)\in [-h,0]\times [z_0,+\infty ) \end{aligned}$$
(21)

and for some wavefront \(\psi _c\), \(b\in {\mathbb {R}}\) and \(q>0\)

$$\begin{aligned} |v_0(s,z)-\psi _c(z)|\le q\eta _{b}(z)\quad \hbox {for all}\quad (s,z)\in [-h,0]\times {\mathbb {R}}, \end{aligned}$$
(22)

then there exists \(C=C({\bar{g}},m_K,b)>0\) and \(\gamma _0\ge 0\) satisfying

$$\begin{aligned} -\lambda _c ^2+c\lambda _c+1\ge \gamma _0+ L_ge^{\gamma _0 h}e^{-\lambda _c ch}\ \hbox {and}\quad L_g({\mathcal {I}}_K)\le e^{-\gamma _0 h}(1-\gamma _0), \end{aligned}$$
(23)

such that

$$\begin{aligned} |v(t,z)-\psi _c(z)|\le Cqe^{-\gamma _0 t},\quad \hbox {for all}\quad (t,z)\in [-h,\infty )\times {\mathbb {R}}. \end{aligned}$$
(24)

Remark 10

(Crossing-monostable case) In [41] Wu et al. established the local stability for sufficiently fast wavefronts of the so-called crossing-monostable case. Theorem 9 generalizes those results by including global perturbations of wavefronts. More precisely, we suppose that for some positive number \(K\ge \kappa \), the birth function g is such that

(\(C_1\)):

g satisfies (M)

(\(C_2\)):

\(g(u)\le {\bar{g}}(u):=\max \{g'(0)u,K\}\) for all \(u\ge 0\) and

(\(C_3\)):

\(L_g({\mathcal {I}}_K)<1,\)

then the non decreasing function \({\bar{g}}(u)\) clearly satisfies \(L_{{\bar{g}}}=g'(0)\le L_g\), therefore g satisfies the conditions of Theorem 9. Hence if g satisfies (\(C_1\))-(\(C_3\)) then (22) implies (24). Note that the condition \((A_3)\) in [41, Theorem 2.4] is essentially our condition (\(C_3\)). Moreover, by (2) and the definition of \({\mathfrak {c}}_*(h)\) we have \({\mathfrak {c}}_*(h)\) is a non-increasing function of h, therefore we have \({\mathfrak {c}}_{*}(h)\le {\mathfrak {c}}_{*}(0)=2\sqrt{L_g-1}\) for all \(h\ge 0\), so that we have improved the minimal speed \({\tilde{c}}:=2\sqrt{2(L_g-1)}\) given in [41] for the local stability of wavefronts with speed \(c>{\tilde{c}}.\)

Now, if we take \({\bar{g}}(u)=\max _{s\in [0,u]}g(s)\) then we have that \(K=M_g:=\max _{s\in [0,\kappa ]}g(s)\) and by writing \(m_g=\min _{u\in [\kappa ,M_g]}g(u)\) and \({\mathcal {I}}_K=I_g:=[m_g, M_g]\) the following global stability result is obtained

Corollary 11

Let g satisfy (M) such that \(L_g (I_g ) < 1\). If \(\psi _c\) is a semi-wavefront with speed \(c > {\mathfrak {c}}_*\), then \(\psi _c\) is globally stable in the sense of Theorem 9.

Corollary 11 generalizes results for wavefronts which assume the D-K condition (see, e.g. [33]). In the Allee case with monotone g, Corollary 11 is an improvement, in terms of the globality of the disturbance, of [33, Theorem 2] for wavefronts with a speed greater than \({\mathfrak {c}}_*\) and it also gives us an exponential convergence rate for these waves. In this regard, exponential (in the time) stability as in (24) for pushed wavefronts was not studied in [32] but a recent work [42] by Wu, Niu and Hsu, has given a positive answer to this problem.

This paper is organized as follows. The linear theorems (Theorems 1 and 2) are proven in Sect. 2. Finally, results on the stability of semi-wavefronts are proven in Sect. 3.

2 Proof of Linear Theorems

In order to demonstrate both Theorems 1 and 2, the following two lemmas will be needed. The first one is an abstract version of the Halanay type inequalities [17]

Lemma 12

(Halanay Type Inequality) Let X be a complex Banach space. Suppose that \(\sigma ,k\in {\mathbb {C}}\) and \(h>0\). If \(r\in C([-h,\infty ),X)\) is a function satisfying:

$$\begin{aligned} r_t(t)= \sigma r(t)+kr(t-h), \quad \hbox {a.e.,} \end{aligned}$$

then

$$\begin{aligned} |r(t)|_{X}\,\le \sup _{s\in [-h,0]} |r(s)|_X \ e^{\max \{0,\lambda \}h }\, e^{\lambda t},\quad \hbox {for all}\quad t>-h, \end{aligned}$$
(25)

where \(\lambda \) is the only real root of the equation :

$$\begin{aligned} \lambda =Re(\sigma )+|k|e^{-\lambda h}. \end{aligned}$$
(26)

Moreover

  1. (i)

    \(\lambda \le 0 \iff -Re(\sigma )\ge |k|. \)

  2. (ii)

    \(\lambda = 0 \iff -Re(\sigma )= |k|. \)

Proof

It is clear that:

$$\begin{aligned} \frac{d}{dt}(r(t)e^{-\sigma t})=ke^{-\sigma t}r(t-h)\quad a.e. \end{aligned}$$

and from here, it is obtained that \(|r(t)|_X\) meets the following inequality:

$$\begin{aligned} x(t)\le |k|\int _{0}^{t} e^{Re(\sigma )(t-s)}x(s-h)ds+x(0)e^{Re(\sigma )t}\quad \hbox {for all}\quad t>0 \end{aligned}$$
(27)

We note that for \(A\in {\mathbb {R}}\) the function \(e_A (t)=Ae^{\lambda t}\) meets (27) with equality. Now, for \(A:= \sup _{s\in [-h,0]}|r(s)|_X\,e^{\max \{0,\lambda \} h}\) the function \(\delta (t)=|r(t)|_X-e_{A}\) satisfies (27) for \(t\in [0,h]\) and therefore \(\delta (t)\le 0\) for all \(t\in [0,h]\). Similarly, it is concluded that \(\delta (t)\le 0\) for the intervals \([h,2h],[2h,3h]\ldots \) This proves (25).

Let us prove (i). If \(-Re(\sigma )\ge |k|\) then: \(\lambda \le |k|(e^{-h\lambda }-1)\) which necessarily implies that \(\lambda \le 0\). Otherwise, if \(\lambda \le 0\) let us suppose that \(-Re(\sigma )<b\), then \(\lambda >|k|(e^{-h\lambda }-1)\) which is a contradiction.

In order to prove (ii) let us note that since the derivative of \(f(\lambda ):=\lambda -Re(\sigma )-|k|e^{-h\lambda }\) is always positive then \(f(\lambda )\) has at most one zero. So, if \(Re(\sigma )=|k|\) then \(\lambda =0\) is the only solution of (26), this proves (ii). \(\square \)

Now, let us consider the function \(\lambda : {\mathbb {R}}\rightarrow {\mathbb {R}}\) defined by

$$\begin{aligned} \lambda (\zeta )=-\zeta ^2 +p+qe^{-h\lambda (\zeta )}, \end{aligned}$$
(28)

where \(q\ge 0\). Next, we proceed to estimate the even function \(\lambda (\zeta )\).

For \(\epsilon _h=\frac{1}{1+h(\gamma -p)}\) we define the function

$$\begin{aligned} \alpha _{h}(\zeta ):=-\frac{1}{h}\log (1+h\epsilon _h \zeta ^2). \end{aligned}$$

Here \(\gamma \in {\mathbb {R}}\) is defined by (13) for any \(p\in {\mathbb {R}}\) and \(q\ge 0\).

Lemma 13

If \(\lambda \) is defined by (28) then

$$\begin{aligned} -\epsilon _h\zeta ^2+\gamma \le \lambda (\zeta )\le \alpha _{h}(\zeta )+\gamma \quad \hbox {for all}\ \zeta \in {\mathbb {R}}. \end{aligned}$$
(29)

Moreover, if \(q>0\) then

$$\begin{aligned} \lim _{|\zeta |\rightarrow \infty }q^{-1}\zeta ^2\,e^{h\lambda (\zeta )}=1. \end{aligned}$$
(30)

Remark 14

The function \(\alpha _{h}\) is a generalization of the function \(\alpha _0 (\zeta ):=-\zeta ^2=\lim _{h\rightarrow 0}\alpha _h(\zeta )\) for each \(\zeta \in {\mathbb {R}}\). Also, when \(h=0\) then \(\gamma =p+q\) [according to definition (13)] therefore \(\lambda (\zeta )=-\zeta ^2+\gamma \) in (28). Thus, by passing the limit \(h\rightarrow 0\) in (29) we have the equality \(-\zeta ^2=\lambda (\zeta )-\gamma =\alpha _0(\zeta ).\) In this regard, the estimates in (29) are sharp.

Proof

Let us denote \(\beta (\zeta )=\lambda (\zeta )-\alpha (\zeta )-\gamma \). Then \(\beta (\zeta )\) satisfies the following equation

$$\begin{aligned} \beta (\zeta )=-\zeta ^{2}+\frac{1}{h}\log (1+h\epsilon _h\zeta ^2)-\gamma +p+qe^{-h\gamma }(1+h\epsilon _h\zeta ^2)e^{-h\beta (\zeta )}. \end{aligned}$$

From Lemma 12 we have that \(\beta (\zeta )\le 0\) if and only if:

$$\begin{aligned} \zeta ^{2}-\frac{1}{h}\log (1+h\epsilon _h\zeta ^2)+\gamma -p\ge qe^{-h\gamma }(1+h\epsilon _h\zeta ^2). \end{aligned}$$
(31)

Now, using \(\log (1+x)\le x\), fo rall \(x\ge 0\), in order to obtain (31) it is enough to have

$$\begin{aligned}&\zeta ^2-\epsilon _{h}\zeta ^2+\gamma -p\ge qe^{-h\gamma }(1+h\epsilon _h\zeta ^2)\quad \hbox {for all}\quad \zeta \in {\mathbb {R}}\\&\quad \iff (1-\epsilon _{h}-qh\epsilon _h e^{-h\gamma })\zeta ^2+\gamma -p- qe^{-h\gamma }\ge 0\quad \hbox {for all}\quad \zeta \in {\mathbb {R}},\\ \end{aligned}$$

which is a consequence of definition of \(\gamma \) and \(\epsilon _h\). So that, this proves the upper estimate in (29)

To complete left hand side of (29) we note that due to (28), (13) and upper estimate in (29)

$$\begin{aligned} \lambda (\zeta )\ge & {} -\zeta ^2+\gamma -q^{-\gamma h}+qe^{-h\gamma }[1+h\epsilon _h\zeta ^2]\\= & {} -\epsilon _h\zeta ^2+\gamma . \end{aligned}$$

Next, by multiplying (28) by \(e^{h\lambda (\zeta )}\) and by using that \(\lambda (\zeta )\rightarrow -\infty \) as \(|\zeta |\rightarrow +\infty \) [which is obtained from upper estimation in (29)] we conclude

$$\begin{aligned} \lim _{\zeta \rightarrow \pm \infty }e^{h\lambda (\zeta )}\zeta ^2=p \end{aligned}$$

which implies (30). \(\square \)

Consider the following equation

$$\begin{aligned} u_t(t,z)=u_{zz}(t,z)+d_1u_z(t,z)+d_2u(t,z)+e^{-\lambda z}g(e^{\lambda (z-ch)}u(t-h,z-ch))\nonumber \\ \end{aligned}$$
(32)

where \(d_1,d_2,\lambda \in {\mathbb {R}}\)

Proposition 15

If \(u_0\in C([-h,0]; L^{\infty }({\mathbb {R}})\cap C^{0,\alpha }({\mathbb {R}}))\), some \(\alpha \in (0,1]\), then there is a unique solution u(tz) of (32) with initial data \(u_0\) and this solution satisfies \(u(\cdot +kh, \cdot )\in C([-h,0]; L^{\infty }({\mathbb {R}})\cap C^{0, \alpha }({\mathbb {R}}))\) for all \(k\in {\mathbb {Z}}_+\). Moreover, if \(u_0\in C([-h,0]; L^1({\mathbb {R}}))\) and \(\lambda =0\) in (32) then \(u(t,\cdot ), u_z(t, \cdot )\in L^{1}({\mathbb {R}})\) for all \(t\ge 0\) and \(u_{zz}(t, \cdot )\in L^1({\mathbb {R}})\) for all \(t>h\).

Proof

By defining

$$\begin{aligned} d_3(t,z):= e^{-\lambda ch}g(e^{\lambda (z-ch)}u(t-h,z-ch))/e^{\lambda (z-ch)}u(t-h, z-ch) \end{aligned}$$

we have \(d_3\in L^{\infty }({\mathbb {R}}_+\times {\mathbb {R}})\) and the function u satisfies

$$\begin{aligned} u_t(t,z)=u_{zz}(t,z)+d_1u_z(t,z)+d_2u(t,z)+d_3(t,z)u(t-h,z-ch) \end{aligned}$$
(33)

By making the change of variables \({\bar{u}}(t,z):=u(t,z-d_1t)e^{-d_2 t}\) the equation (33) is reduced to an inhomogeneous heat equation

$$\begin{aligned} {\bar{u}}_t(t,z)={\bar{u}}_{zz}(t,z)+f(t,z), \end{aligned}$$
(34)

where

$$\begin{aligned} f(t,z)=e^{-d_2h}d_3(t,z-d_1t){\bar{u}}(t-h,z-h(c+d_1)), \end{aligned}$$
(35)

Now, note that for \(1\le p\le \infty \)

$$\begin{aligned} |f(t,\cdot )|_{L^{p}({\mathbb {R}})}\le e^{-d_2h}|d_3|_{L^{\infty }}\max _{s\in [-h,0]}|{\bar{u}}_0(s, \cdot )|_{L^{p}({\mathbb {R}})}\quad \hbox {for all}\ t\in [0,h]. \end{aligned}$$
(36)

Similarly, by using the definition of \(d_3\), we get

$$\begin{aligned} |f(t,\cdot )|_{C^{0,\alpha }({\mathbb {R}})}\le L_g\,e^{-d_2h} \max _{s\in [-h,0]}|{\bar{u}}_0(s, \cdot )|_{C^{0,\alpha }({\mathbb {R}})} \quad \hbox {for all} \ t\in [0,h]. \end{aligned}$$
(37)

So that, by [14, Chapter 1, Theorems 12 and 16] there exist a unique solution to (32) and this solution satisfies

$$\begin{aligned} {\bar{u}}(t):= \Gamma _t*{\bar{u}}(0)+\int _0^t\Gamma _{t-s}*f(s)ds, \end{aligned}$$
(38)

where \(\Gamma _t\) is the one-dimensional heat kernel.

Now, we take \(1\le p\le \infty \). Then, for \(t\in [0,h]\) and \(t_n\rightarrow t\) we have

$$\begin{aligned} \nonumber |{\bar{u}}(t)-{\bar{u}}(t_n)|_{L^p}\le & {} |\Gamma _t-\Gamma _{t_n}|_{L^1}\,|{\bar{u}}(0)|_{L^p}+\int _{0}^t|\Gamma _{t-s}-\Gamma _{t_n-s}|_{L^1}|f(s)|_{L^p}ds\\&+ \int _{t}^{t_n}|\Gamma _{t_n-s}|_{L^1}|f(s)|_{L^p}ds, \end{aligned}$$
(39)

and by using (36),

$$\begin{aligned}&|{\bar{u}}(t)-{\bar{u}}(t_n)|_{L^p}\le (|\Gamma _t-\Gamma _{t_n}|_{L^1}\nonumber \\&\quad +\int _{0}^t|\Gamma _{t-s}-\Gamma _{t_n-s}|_{L^1}ds+| t-t_n|)R \max _{s\in [-h,0]}|{\bar{u}}_0(s,\cdot )|_{L^p({\mathbb {R}})}, \end{aligned}$$
(40)

where \(R=\max \{1, e^{-d_2h}|d_3|_{L^{\infty }}\}\). Since \(|\Gamma _{t_n}|_{L^1}=|\Gamma _t |_{L^1}=1\) the last inequality implies \(|{\bar{u}}(t)-{\bar{u}}(t_n)|\rightarrow 0\) as \(t_n\rightarrow t\), therefore if \(u_0(\cdot , \cdot )\in C([-h,0]; L^p({\mathbb {R}}))\) then \(u(\cdot +h, \cdot )\in C([-h,0]; L^p({\mathbb {R}}))\). Similarly, we get \(u(\cdot +h, \cdot )\in C([-h,0]; C^{0, \alpha }({\mathbb {R}}))\) whenever \(u_0(\cdot , \cdot )\in C([-h,0]; C^{0,\alpha }({\mathbb {R}}))\)

Analogously, by using the initial data \(u(t+h,\cdot ), u(t+2h,\cdot )\ldots \) we obtain \(u(\cdot +kh, \cdot )\in C([-h,0]; L^p({\mathbb {R}})\cap C^{0,\alpha }({\mathbb {R}}))\) for \(k=2,3\ldots \) Therefore, with \(p=\infty \) we obtain the first assertion of the Proposition 15.

Otherwise, if \(u_0\in C([-h,0], L^1({\mathbb {R}}))\) then with \(p=1\) we get \(u(\cdot + kh, \cdot )\in C([-h,0]; L^1({\mathbb {R}}))\) for all \(k\in {\mathbb {Z}}_+\). Then, note that by (38) for \(t>0\) we get

$$\begin{aligned} {\bar{u}}_z(t,z)= & {} \int _{{\mathbb {R}}}\frac{(y-z)e^{-(z-y)^2/4t}}{4t^{3/2}\sqrt{\pi }}u_0(0, y)dy\nonumber \\&+ \int _{0}^t\int _{{\mathbb {R}}}\frac{(y-z)e^{-(z-y)^2/4(t-s)}}{4(t-s)^{3/2}\sqrt{\pi }}f(s,y)dyds, \end{aligned}$$
(41)

and using (36) with \(p=1\) for \(t\in (0,h]\) we obtain

$$\begin{aligned} |{\bar{u}}_z(t,\cdot )|_{L^1({\mathbb {R}})}\le & {} \frac{|u_0(0,\cdot )|_{L^1({\mathbb {R}})}}{\sqrt{\pi t}}\int _{{\mathbb {R}}}|y|e^{-y^2}dy\nonumber \\&+2\sqrt{\frac{t}{\pi }}\int _{{\mathbb {R}}}|y|e^{-y^2}dy \max _{s\in [-h,0]}|f(s,\cdot )|_{L^1({\mathbb {R}})}\nonumber \\\le & {} (\frac{1}{\sqrt{t}}+2\sqrt{t}e^{-d_2h}|d_3|_{L^{\infty }({\mathbb {R}}_+\times {\mathbb {R}})})\, \frac{1}{\sqrt{\pi }}\max _{s\in [-h,0]}|{\bar{u}}_0(s, \cdot )|_{L^1({\mathbb {R}})}\nonumber \\ \end{aligned}$$
(42)

and by using the initial data \({\bar{u}}(t+h,\cdot ), {\bar{u}}(t+2h,\cdot )\ldots \), with \(t\in (0,h]\), we obtain \({\bar{u}}_z(t+kh, \cdot )\in L^1({\mathbb {R}})\) for \(k\in {\mathbb {Z}}_+\) and \(t\in (0,h]\). Moreover, if we differentiate in (32) and proceed as in (39) and (40) then we have \(|{\bar{u}}_z(t, \cdot )|_{L^1({\mathbb {R}})}\) continuously depends on \(t\in {\mathbb {R}}_+\).

Finally, if \(T>h\) then \({\bar{u}}(T+\cdot , \cdot )\in C([-h,0]; L^1({\mathbb {R}}))\), by taking \(\lambda =0\), we obtain \({\bar{u}}_z(t,z)\) satisfies (33) with \(d_3(t,z)= g'({\bar{u}}(t-h, z-ch))\) and taking as initial datum the function \({\bar{u}}(T+s, z)\) and using (42) (replacing \({\bar{u}}_z\) by \({\bar{u}}_{zz}\)) we obtain \({\bar{u}}_{zz}(T+t, \cdot )\in L^{1}({\mathbb {R}})\) for all \(t\in (0,h]\). Similarly, by using the initial data \({\bar{u}}_z(t+h,\cdot ), {\bar{u}}_z(t+2h,\cdot )\ldots \), with \(t\in (0,h]\), we obtain \({\bar{u}}_{zz}(t+T+kh, \cdot )\in L^1({\mathbb {R}})\) for \(k\in {\mathbb {Z}}_+\) and \(t\in (0,h]\), which completes the proof. \(\square \)

Remark 16

Since by Proposition 15 \(u(\cdot +hk, \cdot )\in C([-h, 0], L^{\infty }({\mathbb {R}}))\) for all \(k\in {\mathbb {Z}}\) then for each \(t>0\) we have \(f(\cdot , \cdot )\in C([0, t], L^{\infty }({\mathbb {R}}))\), therefore

$$\begin{aligned} \left| \int _{{\mathbb {R}}}\frac{(y-z)e^{-(z-y)^2/4(t-s)}}{4(t-s)^{3/2}\sqrt{\pi }}f(s,y)dy\right| \le \frac{C}{\sqrt{t-s}}\quad \hbox {for all}\quad t>s, \end{aligned}$$

for some constant \(C>0\) (which does not depend on (zts)) so that from (41) we conclude \(u(t, \cdot )\in C^1({\mathbb {R}})\) for all \(t>0\).

Proof of Theorem 1

By using Proposition 15 with \(d_1=m\), \(d_2=p\), \(\lambda =0\) and \(g(u)=qu\) we get \(u(t,\cdot ), u_z(t, \cdot ), u_{zz}(t, \cdot )\in L^1({\mathbb {R}})\) for all \(t>h\). Next, by applying the Fourier transform, here

$$\begin{aligned} {\hat{u}}(z):= \frac{1}{2\pi }\int _{{\mathbb {R}}} e^{-izy}u(y)dy, \end{aligned}$$

to Eq. (12) we have

$$\begin{aligned} {\hat{u}}_t(t,\zeta )=\sigma (\zeta ) {\hat{u}}(t,\zeta )+k(\zeta ){\hat{u}}(t-h,\zeta )\quad \hbox {for all}\quad t>h, \end{aligned}$$

where \(\sigma (\zeta )=-\zeta ^2+im\zeta +p\) and \(k(\zeta )=qe^{-id\zeta }.\)

Since \(-Re(\sigma (\zeta ))\ge |k(\zeta )|\), by Lemma 12 we obtain \(\lambda (\zeta )\le 0\) for all \(\zeta \in {\mathbb {R}}\) and :

$$\begin{aligned} |{\hat{u}}(t,\zeta )|\le C_{u_0} e^{\lambda (\zeta )t}\quad \hbox {for all} \ \zeta \in {\mathbb {R}}. \end{aligned}$$

If \(t>h\) then by the Fourier’s inversion formula (since by Remark 16 \(u(t,\cdot )\in C^1({\mathbb {R}})\) for \(t>0\)) and Lemma 13, we have

$$\begin{aligned} |u(t,x)|\le \frac{1}{2\pi }\int _{{\mathbb {R}}} |{\hat{u}}(t,\zeta )|d\zeta \le \frac{C_{u_0}}{2\pi }\int _{{\mathbb {R}}} e^{\lambda (\zeta )t}d\zeta \le \frac{C_{u_0}}{2\pi }e^{\gamma t}\int _{{\mathbb {R}}} \frac{d\zeta }{(1+\epsilon \zeta ^2)^{\frac{t}{h}}}. \end{aligned}$$

Moreover, by Bernoulli’s inequality, we conclude that

$$\begin{aligned} \int _{{\mathbb {R}}} \frac{d\zeta }{(1+\epsilon \zeta ^2)^{\frac{t}{h}}}\le \int _{{\mathbb {R}}}\frac{d\zeta }{1+\frac{t\epsilon }{h}\zeta ^2}=\frac{1}{\sqrt{t}}\left[ \sqrt{\frac{h}{\epsilon }}\int _{{\mathbb {R}}}\frac{d\zeta }{1+\zeta ^2}\right] =\frac{1}{\sqrt{t}}\sqrt{\frac{h}{\epsilon }}\pi . \end{aligned}$$

\(\square \)

Proof of Theorem 2

If we make the change of variable \(v(t,x)=e^{\frac{m}{2}x}u(t,x)\), then v(tx) solves

$$\begin{aligned} v_t(t,x)=v_{xx}(t,x)+\left( p-\frac{m^2}{4}\right) v(t,x) +qv(t-h,x). \end{aligned}$$
(43)

By applying the Fourier transform to (43) we get

$$\begin{aligned} {\hat{v}}_t(t,z)=\left( -z^2+p-\frac{m^2}{4}\right) {\hat{v}}(t,z)+q{\hat{v}}(t-h,z)\quad \hbox {for all}\quad t>0 \end{aligned}$$
(44)

Let us note that due to \(q \ge 0\), we have that (44) satisfies the Comparison Principle; that is, if for each \(z\in {\mathbb {R}}\) we consider two solutions v(s) and w(s) of (44) defined on \([-h, +\infty )\) then, by denoting \(\mathfrak {R}({\hat{u}}(t))=u_1(t), \mathfrak {I}({\hat{u}}(t))=u_2(t)\), the inequality

$$\begin{aligned} v_i(s)\le w_i(s)\quad \hbox {for all} \ s\in [-h,0] \ \hbox {and}\quad i=1,2. \end{aligned}$$

implies

$$\begin{aligned} v_i(s)\le w_i(s)\quad \hbox {for all} \ s\in [-h,+\infty ) \ \hbox {and}\quad i=1,2 \end{aligned}$$

Let us denote by \(e_{A}(t,z)=Ae^{\lambda (z)t}\), where \(\lambda (z)\) satisfies

$$\begin{aligned} \lambda (z)=-z^2+p-\frac{m^2}{4}+qe^{-\lambda (z)h}. \end{aligned}$$
(45)

Let us note that \(e_A(t,z)\) satisfies (44) for all \(A\in {\mathbb {C}}\). Also, let us denote that

$$\begin{aligned} m_i(z)= & {} \min _{s\in [-h,0]}(v_i(s,z) e^{-\lambda (z)s})\quad \hbox {and}\quad \\ M_i(z)= & {} \max _{s\in [-h,0]}(v_i(s,z) e^{-\lambda (z)s})\quad i=1, 2 \end{aligned}$$

then we have that

$$\begin{aligned} e_{m_i}(s,z)\le v_i(s,z)\le e_{M_i}(s,z)\quad \hbox {for all} \ (s,z)\in [-h,0]\times {\mathbb {R}}; i=1,2. \end{aligned}$$

By the comparison principle applied to real and imaginary part in (44), we have that

$$\begin{aligned} e_{m_i}(t,z)\le v_i(t,z)\le e_{M_i}(t,z)\quad \hbox {for all}\quad (t,z)\in [-h,\infty )\times {\mathbb {R}}; i=1,2\nonumber \\ \end{aligned}$$
(46)

or

$$\begin{aligned} m_i(z)e^{\lambda (z)t}\le v_i(t,z)\le M_i(z)e^{\lambda (z)t}\quad \hbox {for all}\quad (t,z)\in [-h,\infty )\times {\mathbb {R}}; i=1,2\nonumber \\ \end{aligned}$$
(47)

Now, by the Fourier inversion formula, we have that

$$\begin{aligned} v(t,x)=\frac{1}{\sqrt{t}}\int _{{\mathbb {R}}}e^{\frac{xy}{\sqrt{t}}}{\hat{v}}(t,y/\sqrt{t})dy. \end{aligned}$$
(48)

However, if we apply Lemma 13 to (45) with \(\gamma =\sigma \) we have

$$\begin{aligned} \lim _{t\rightarrow \infty }t[\lambda (y/\sqrt{t})-\sigma ]=-\frac{y^2}{1+hqe^{-\sigma h}}. \end{aligned}$$
(49)

and due to \(v(s, \cdot ) \in L^1 ({\mathbb {R}})\) by the Lebesgue’s dominated convergence theorem

$$\begin{aligned} \lim _{t\rightarrow \infty }M_1(y/\sqrt{t})=\lim _{t\rightarrow \infty }m_1(y/\sqrt{t})=\int _{{\mathbb {R}}}e^{\frac{m}{2}x}v(s,x)dx \end{aligned}$$

and

$$\begin{aligned} \lim _{t\rightarrow \infty }M_2(y/\sqrt{t})=\lim _{t\rightarrow \infty }m_2(y/\sqrt{t})=0. \end{aligned}$$

Therefore by (47)

$$\begin{aligned} \lim _{t\rightarrow \infty }{\hat{v}}(t, y/\sqrt{t})=e^{-\frac{y^2}{1+hqe^{-\sigma h}}}\int _{{\mathbb {R}}}e^{\frac{m}{2}x}v(s,x)dx \end{aligned}$$
(50)

However, by (25) there exists \(C (p, q, m) > 0\) such that

$$\begin{aligned} |{\hat{v}}(t,y/\sqrt{t})|\le C\sup _{s\in [-h,0]}|{\hat{v}}(s,y/\sqrt{t})| e^{\lambda (y/\sqrt{t})t} \end{aligned}$$

but by (29) and Bernoulli’s Inequality

$$\begin{aligned} |{\hat{v}}(t,y/\sqrt{t})|\le \frac{Ce^{|\sigma |h}||e^{\frac{m}{2}\cdot }u_0(\cdot )||_{L^1({\mathbb {R}})}}{1+\epsilon _h y^2}\quad \hbox {for all}\quad t>0. \end{aligned}$$
(51)

Finally, by (48), (51), Lebesgue’s dominated convergence theorem and (50), the result obtained. \(\square \)

3 Proof of results of stability of semi-wavefronts

Proof of Theorem 3

The first assertion follows from Proposition 15 with \(\lambda =0\). Next, for a solution w(tz) of (17), let us denote the function \({\tilde{w}}(t,z)=\xi _c(z)w(t,z)\) which satisfies

$$\begin{aligned} {\tilde{w}}_t(t,z)= & {} {\tilde{w}}_{zz}(t,z)+m{\tilde{w}}_z(t,z)+p{\tilde{w}}(t,z)\\&+\xi _c(z)g(\xi _c(-z+ch){\tilde{w}}(t-h,z-ch)). \end{aligned}$$

We consider the linear operator

$$\begin{aligned} {\mathcal {L}}\delta (t,z):=\delta _{zz}(t,z)+m\delta _z(t,z)+p\delta (t,z)-\delta _t(t,z). \end{aligned}$$

If \(\delta _{\pm }(t,z):=\pm [{\tilde{v}}(t,z)-{\tilde{\psi }}(t,z)]-u(t,z)\), then by (18): \(\delta _{\pm }(s,z)\le 0\) for \((s,z)\in [-h,0]\times {\mathbb {R}}\). For \((t,z)\in [0,h]\times {\mathbb {R}}\) by (17) and (18) we have

$$\begin{aligned} {\mathcal {L}}\delta _{\pm }(t,z)= & {} \mp \xi (z)[g(\xi (-z+ch){\tilde{\psi }}(t-h,z-ch))\\&-g(\xi (-z+ch){\tilde{v}}(t-h,z-ch))]-{\mathcal {L}}u(t,z)\\\ge & {} -L_ge^{-\lambda ch}|{\tilde{v}}(t-h,z-ch)-{\tilde{\psi }}(t-h,z-ch)|-{\mathcal {L}}u(t,z)\\\ge & {} - L_ge^{-\lambda ch}u(t-h,z-ch)-{\mathcal {L}}u(t,z)= 0. \end{aligned}$$

Now, by Proposition 15, \({\tilde{w}}(\cdot +kh,\cdot )\in C([-h,0]; L^{\infty }({\mathbb {R}}))\) for all \(k\in {\mathbb {Z}}_+\) therefore by using the Phragmèn–Lindelöf principle from [25, Chapter 3, Theorem 1], we have \(\delta _{\pm }(t,z)\le 0\) for \((t,z)\in [0,h]\times {\mathbb {R}}\). The argument is repeated for intervals \([h, 2h], [2h, 3h]\ldots \) to conclude (19). Finally, the estimate in (20) is obtained using Theorem 1. \(\square \)

Remark 17

Note that in Proof of Theorem 3 it was only necessary to have an initial datum \(u_0\) exponentially bounded to apply the Phragmèn–Lindelöf principle in order to obtain estimate (20). So, we could use the elementary exponential solutions of (12) of the form \(u(t, z)=Be^{\gamma t+r\,z}\), with r and \(\gamma \) satisfying

$$\begin{aligned} q(\lambda _c)e^{-rch }\,e^{-\gamma h}=-r^2-(2 \lambda _c-c)r-p(\lambda _c)+\gamma . \end{aligned}$$
(52)

Here, \(\gamma \le 0\) if and only if

$$\begin{aligned} -r^2-(2\lambda _c-c)r-p(\lambda _c)\ge q(\lambda _c)\,e^{-r ch }, \end{aligned}$$
(53)

with \(\gamma =0\) if and only if (53) holds with equality. Thus, for \(c>{\mathfrak {c}}_*\) and \(\lambda _c\in (\lambda _1(c), \lambda _2(c))\) we have \(-p(\lambda _c)>q(\lambda _c)\) and therefore by taking \(r=0\) in (53) we obtain \(\gamma <0\) in (52) and therefore the asymptotic stability of non-critical is obtained. However, when \(c={\mathfrak {c}}_*\) we have \(\lambda _{{\mathfrak {c}}_*}=\lambda _1({\mathfrak {c}}_*)=\lambda _2({\mathfrak {c}}_*)\) and \(-p(\lambda _{{\mathfrak {c}}_*})=q(\lambda _{{\mathfrak {c}}_*})\) in (53), also due to the curves \(-\lambda ^2+c\lambda +1\) and \(L_g\,e^{-\lambda ch}\) in (2) are tangent at \(\lambda =\lambda _{{\mathfrak {c}}_*}\) the function \(\Theta (r):=q(\lambda _{{\mathfrak {c}}_*})\,e^{-r {\mathfrak {c}}_* h }+r^2+(2\lambda _{{\mathfrak {c}}_*}-{\mathfrak {c}}_*)r+p(\lambda _{{\mathfrak {c}}_*})\) holds \(\Theta \, '(0)=0\). Consequently, since \(\Theta \) is strictly convex and \(\Theta (0)=0\) we conclude \(r=0\) is the only solution in (53) and therefore \(-p(\lambda _{{\mathfrak {c}}_*})=q(\lambda _{{\mathfrak {c}}_*})\) implies \(\gamma =0\) in (52). Thus, this approach does not allow us to obtain the asymptotic stability of critical semi-wavefronts.

Theorem 18

Let v(tz) and \(\psi (t,z)\) be solutions of equation (17) for \(c\ge {\mathfrak {c}}_*\). Assume that for some compact interval \(I \subset {\mathbb {R}}\), such that \(L_g (I ) < 1,\) and \(b\in {\mathbb {R}}\) we have

$$\begin{aligned} \psi (t,z),v(t,z)\in I \quad \hbox {for all} \ (t,z)\in [-h,\infty )\times [b-ch,\infty ), \end{aligned}$$
(54)

and for some \(q > 0\) and \(\lambda _c\in [\lambda _1(c), \lambda _2(c)]\)

$$\begin{aligned} |v_0(s,z)-\psi _0(s,z)|\le q\eta _b(z)\quad \hbox {for all}\ (s,z)\in [-h,0]\times {\mathbb {R}}. \end{aligned}$$
(55)

If \(\gamma _0 \ge 0\) satisfies

$$\begin{aligned} -\lambda _c ^2+c\lambda _c+1\ge \gamma _0+ L_ge^{\gamma _0 h}e^{-\lambda _c ch}\ \hbox {and}\quad L_g(I)\le e^{-\gamma _0 h}(1-\gamma _0), \end{aligned}$$
(56)

then

$$\begin{aligned} |v(t,z)-\psi (t,z)|\le qe^{-\gamma _0 t}\eta _b(z)\quad \hbox {for all}\ (t,z)\in [-h,\infty )\times {\mathbb {R}}. \end{aligned}$$
(57)

Proof

We define \(\eta (t,z)=qe^{-\gamma _0 t}\eta _b(z)\) and write the operator

$$\begin{aligned} {\mathcal {L}}_0\delta (t,z):=\delta _{zz}(t,z)-c\delta _z(t,z)-\delta (t,z)-\delta _t(t,z). \end{aligned}$$

Note that by (55) if \(\delta _{\pm }(t,z):=\pm [v(t,z)-\psi (t,z)]-\eta (t,z)\) then \(\delta _{\pm }(s,z)\le 0\) for \((s,z)\in [-h,0]\times {\mathbb {R}}\). Now, for \((t,z)\in [0,h]\times (-\infty ,b]\) due to (17), (22) and (56) we have that

$$\begin{aligned} {\mathcal {L}}_0\delta _{\pm }(t,z)= & {} \pm [-g(v(t-h,z-ch))+g(\psi (t-h,z-ch))]-{\mathcal {L}}_0\eta (t,z)\\\ge & {} qe^{-\gamma _0 t+\lambda _c (z-b)}[-L_ge^{\gamma _0 h}e^{-\lambda ch}-(\lambda _c^{2}-c\lambda _c-1+\gamma _0)]\ge 0. \end{aligned}$$

Similarly, if \((t,z)\in [0,h]\times [b,\infty )\) we obtain:

$$\begin{aligned} {\mathcal {L}}_0\delta _{\pm }(t,z)= & {} \pm [-g(v(t-h,z-ch))+g(\psi (t-h,z-ch))]-{\mathcal {L}}_0\eta (t,z)\\\ge & {} qe^{-\gamma _0 t}[-L_g(I)e^{\gamma _0 h}\eta (z-ch)-(-1+\gamma _0)]\\\ge & {} qe^{-\gamma _0 t}[-L_g(I)e^{\gamma _0 h}+1-\gamma _0]\ge 0. \end{aligned}$$

Now, as in the proof of the [33, Lemma 1], due to

$$\begin{aligned} \frac{\partial \delta _{\pm }(t,b+)}{\partial z}- \frac{\partial \delta _{\pm }(t, b-)}{\partial z} > 0, \end{aligned}$$
(58)

we have that \(\delta _{\pm }(t,z) \le 0\) for all \(t \in [0,h], \ z \in {\mathbb {R}}\). Indeed, otherwise there exists \(r_0> 0\) such that \(\delta (t,z)\) restricted to any rectangle \(\Pi _r= [-r,r]\times [0,h]\) with \(r>r_0\), reaches its maximum positive value \(M_r >0\) at some point \((t',z') \in \Pi _r\).

We claim that \((t',z')\) belongs to the parabolic boundary \(\partial \Pi _r\) of \(\Pi _r\). Indeed, suppose on the contrary, that \(\delta (t,z)\) reaches its maximum positive value at some point \((t',z')\) of \(\Pi _r\setminus \partial \Pi _r\). Then clearly \(z' \not =z_*\) because of (58). Suppose, for instance that \(z' > z_*\). Then \(\delta (t,z)\) considered on the subrectangle \(\Pi = [z_*,r]\times [0,h]\) reaches its maximum positive value \(M_r\) at the point \((t',z') \in \Pi \setminus \partial \Pi \). Then the classical results [25, Chapter 3, Theorems 5,7] show that \(\delta (t,z) \equiv M_r >0\) in \(\Pi \), a contradiction.

Hence, the usual maximum principle holds for each \(\Pi _r, \ r \ge r_0,\) so that we can appeal to the proof of the Phragmèn–Lindelöf principle from [25] (see Theorem 10 in Chapter 3 of this book), in order to conclude that \(\delta (t,z) \le 0\) for all \(t \in [0,h], \ z \in {\mathbb {R}}\).

We can again repeat the above argument on the intervals [h, 2h],  \([2h, 3h], \dots \) establishing that the inequality \(w_-(t,z) \le w(t,z)\le w_+(t,z),\) \(z\in {\mathbb {R}},\) holds for all \(t \ge -h\). \(\square \)

Remark 19

We can generalize the function \(\eta _b (z)\) for \(b = +\infty \) and, thus, have \(\eta _{\infty }(z)=\xi _c(-z)\). In this proof, it was not necessary to use the condition (54) for \(z \le b\) so by replacing \(\xi _c(-z)\) by \(\eta _b(z)\) it can be concluded that (55) implies (57).

Corollary 20

(Local stability) Suppose that there exist \(M, b \in {\mathbb {R}}\) and \(l_0 > 0\), such that:

$$\begin{aligned} \psi (t,z)\in [M-l_0,M+l_0]\quad \hbox {for all}\quad (t,z)\in [-h,\infty )\times [b-ch,\infty ), \end{aligned}$$
(59)

and that for some \(l_1 > l_0\) the initial data satisfy

$$\begin{aligned} |v_0(s,z)-\psi _0(s,z)|<(l_1-l_0)e^{-\gamma _0 s}\eta _b(z)\quad \hbox {for all} \quad (s,z)\in [-h,0]\times {\mathbb {R}}.\nonumber \\ \end{aligned}$$
(60)

where \(\gamma _0\ge 0\) is defined by (56). If \(L_g({\mathcal {I}}_1) < 1\), where \({\mathcal {I}}_1 := [M - l_1, M + l_1]\), then

$$\begin{aligned} |v(t,z)-\psi (t,z)|\le (l_1-l_0)e^{-\gamma _0 t}\eta _b(z)\quad \hbox {for all} \ (t,z)\in [-h,\infty )\times {\mathbb {R}}. \end{aligned}$$
(61)

Proof

Clearly, \(\psi (t,z)\in {\mathcal {I}}_1\) for all \( (t,z)\in [-h,\infty )\times [b-ch,\infty ).\) Now if we suppose that the inequality in (60) is satisfied for \(v_0(s, z)= v(hk+s, z)\) and \(\psi _0(s, z)=\psi (hk+s, z)\), with \(k\in {\mathbb {Z}}_+\), then \(v(hk+t-h, z)\in {\mathcal {I}}_1\) for all \((t, z)\in [0, h]\times {\mathbb {R}}\) and, arguing as in the proof of Theorem 18, we get

$$\begin{aligned} {\mathcal {L}}_0\,\delta _{\pm }^k(t,z)\le 0\quad \hbox {for all}\quad (t, z)\in [0, h]\times {\mathbb {R}}, \end{aligned}$$

where \(\delta _{\pm }^k(t, z)=\pm [v(hk+t,z)-\psi (hk+t,z)]- (l_1-l_0)e^{-\gamma _0 (hk+t)}\eta _b(z)\) and from [33, Lemma 1] we conclude

$$\begin{aligned} \delta _{\pm }^k(t,z)\le 0\quad \hbox {for all}\ (t, z)\in [0, h]\times {\mathbb {R}}. \end{aligned}$$
(62)

But (62) implies \(v((k+1)h+t-h, z)\in {\mathcal {I}}_1\, \hbox {for all} \ (t, z)\in [0, h]\times {\mathbb {R}}\) and, arguing as above, by using (62) we obtain \(\delta _{\pm }^{k+1}(t,z)\le 0\, \hbox {for all}\ (t, z)\in [0, h]\times {\mathbb {R}}.\) Therefore, it is sufficient to suppose (60) in order to conclude (61) for \((t,z)\in [0,h]\times {\mathbb {R}}\) and then we procede inductively to obtain (61) for all \((t,z)\in [-h, \infty )\times {\mathbb {R}}\). \(\square \)

To prove Theorem 9, we will use the following lemma

Lemma 21

Suppose that functions \(g_1,g_2: D\subset {\mathbb {R}}_+\rightarrow {\mathbb {R}}_+\) satisfy: \(g_1(u)\le g_2(u)\) for all \(u\in D\). Let \(v_1(t, z), v_2(t, z) : [-h, \infty )\times {\mathbb {R}}\rightarrow D\) be solutions to (17), with \(g = g_1\) and \(g = g_2\), respectively, such that: \(v_1(s, z) \le v_2 (s, z)\) for \((s, z) \in [h, 0] \times {\mathbb {R}}\). If \(g_1\) or \(g_2\) is a non decreasing function, then we have : \(v_1(t, z) \le v_2(t, z)\) for all \((t, z) \in {\mathbb {R}}_+ \times {\mathbb {R}}.\)

Proof

We take \(\delta (t,z)=v_1(t,z)-v_2(t,z)\). Let us note that if \((t,z)\in [0,h]\times {\mathbb {R}}\) then

$$\begin{aligned} {\mathcal {L}}_0\delta (t,z)=g_2(v_2(t-h,z-ch))-g_1(v_1(t-h,z-ch))\ge 0, \end{aligned}$$

because if \(g_2\) is a non decreasing function we have that

$$\begin{aligned}&g_2(v_2(t-h,z-ch))-g_1(v_1(t-h,z-ch))\\&\quad \ge g_2(v_1(t-h,z-ch))-g_1(v_1(t-h,z-ch))\ge 0, \end{aligned}$$

or if \(g_1\) is a non decreasing function, we have

$$\begin{aligned}&g_2(v_2(t-h,z-ch))-g_1(v_1(t-h,z-ch))\\&\quad \ge g_2(v_2(t-h,z-ch))-g_1(v_2(t-h,z-ch))\ge 0 \end{aligned}$$

Now, as \(\delta (t,z)\le 0\) for all \((t,z)\in [-h,0]\times {\mathbb {R}}\) the Phragmèn–Lindelöf principle from [25][Chapter 3, Theorem 10] implies that \(\delta (t,z)\le 0\) for \((t,z)\in [0,h]\times {\mathbb {R}}\). The argument is repeated for intervals \([h,2h],[2h,3h]\ldots \) \(\square \)

Proof Theorem 9

Let us take \(\epsilon >0\) such that \(L_g({\mathcal {I}}_{\epsilon })<1\), where \({\mathcal {I}}_{\epsilon }:=[m_K-\epsilon ,K+\epsilon ]\subset {\mathbb {R}}_+.\)Then, there is an increasing function \({\bar{g}}_{\epsilon }\) satisfying (M) with positive equilibrium \(\kappa _+\in (K,K+\epsilon )\), \({\mathfrak {c}}_*(L_{{\bar{g}}_{\epsilon }})\le {\mathfrak {c}}_*( L_g)\) and \(g\le {\bar{g}}_{\epsilon }\). Furthermore, there is also an increasing \({\underline{g}}_{\epsilon }\) function meeting (M) with positive equilibrium \(\kappa _-\in (m_K-\epsilon ,m_K)\) and \({\mathfrak {c}}_*(L_{{\underline{g}}_{\epsilon }})\le {\mathfrak {c}}_* (L_g)\) such that: \({\underline{g}}_{\epsilon }(x)\le g(x)\) for \(x\in [0,K+\epsilon ]\).

Now, if \({\bar{v}}(t)\) is the homogenous solution of (17) replacing g by \({\bar{g}}_\epsilon \) with initial datum \(v_0(s)=q, s\in [-h,0],\) and \(c>{\mathfrak {c}}_*(L_g)\) then by Lemma 21 and the global stability of \(\kappa _+\) there is a number \(T>0\) such that

$$\begin{aligned} v(t,z)\le {\bar{v}}(t)\le K+\epsilon \quad \hbox {for all}\quad (t,z)\in [T,+\infty )\times {\mathbb {R}}. \end{aligned}$$
(63)

Next, by (63)

$$\begin{aligned} \sup _{(t,z)\in {\mathbb {R}}_+\times {\mathbb {R}}}v(t,z)=: v_{\infty }<\infty , \end{aligned}$$

and by denoting \(\kappa _*:=\min _{u\in [\kappa ,u_{\infty }]}g(u)\) we take an increasing function \(g_{\infty }\) satisfying (M) with equilibrium \(\kappa _{\infty }\in (\kappa _*-\epsilon ,\kappa _*)\), \({\mathfrak {c}}_*(L_{g_{\infty }})\le {\mathfrak {c}}_*(L_g)\) and \(g_{\infty }(u)\le g(u)\) for all \(u\in [0,u_{\infty }].\) Next, without loss of generality we take \(q_{ \infty }:=\kappa _{\infty }-q_0>0\). Then, by [1, Theorem 3], (21) and (22) there exist a monotone wavefront \(\phi ^{\infty }_c\) to (17) (with nonlinearity \(g_{\infty }\)) such that

$$\begin{aligned} \phi ^{\infty }_c(z)-q\eta _b(z)\le v_0(s,z)\quad (s,z)\in [-h,0]\times {\mathbb {R}}\end{aligned}$$
(64)

thus by [32, Lemma 2.1] there are \(C_0>0\) and \(\gamma \ge 0\) such that

$$\begin{aligned} \phi _c^{\infty }(z-C_0q)-q^{-\gamma t}\eta _b(z)\le v_{\infty }(t,z)\quad (t,z)\times [-h,\infty )\times {\mathbb {R}}\end{aligned}$$
(65)

where \(v_{\infty }(t,z)\) is the solution generated by the initial datum v(sz) to Eq. (17). Now, by applying Lemma 21 with \(D=[0,u_{\infty }]\)

$$\begin{aligned} \phi _c^{\infty }(z-C_0q)-q^{-\gamma t}\eta _b(z)\le v(t,z)\quad (t,z)\times [-h,\infty )\times {\mathbb {R}}. \end{aligned}$$

So, there are \(z_0'\) and \(q'_0>0\) such that

$$\begin{aligned} v(t,z)\ge q'_0>0\quad (t,z)\in [-h,\infty )\times [z'_0,\infty ) \end{aligned}$$
(66)

Otherwise, denoting \({\underline{v}}(t,z)\) the solution of (17) replacing g by \({\underline{g}}_{\epsilon }\) with initial data \({\underline{v}}_0(s,z)=v(s+T+h,z).\) Due to (66) and Remark 19 the initial datum \({\underline{v}}_0\) satisfies (21) and (22). Next, if we denote by \({\underline{v}}(t)\) the homogenous solution of (17) replacing g by \({\underline{g}}_\epsilon \) with initial datum \(v_0(s)=K+\epsilon , s\in [-h,0],\) then by [30, Corollary 2.2,p.82] \({\underline{v}}(t)\) converges monotonically to \(\kappa _-\), therefore

$$\begin{aligned} {\underline{v}}(t,z)\le {\underline{v}}(t)\le K+\epsilon \quad \hbox {for all} \ (t,z)\in [-h,+\infty )\times {\mathbb {R}}\end{aligned}$$

So, for \(c>c(L_g)\) by Lemma 21 (with \(D=[0,K+\epsilon ]\)), Proposition 7 and [33, Theorem 1] there is a wavefront \({\underline{\phi }}_c\) and \(T_0>0\) such that

$$\begin{aligned} m_K-\epsilon \le {\underline{\phi }}_c(z)+\epsilon /2\le {\underline{v}}(t,z)\le v(t,z)\quad \hbox {for all}\quad (t,z)\in [T_0,\infty )^2. \end{aligned}$$
(67)

Thus there is \(T_{v}\) such that the function \({\tilde{v}}(t,z):=v(t+T_v+h,z)\) satisfies (54) with \(b=t_v+ch\) and \(I={\mathcal {I}}_{\epsilon }\). Analogously, for some \(T_{\psi _c}\) we have \(\psi _c(z)\in {\mathcal {I}}_{\epsilon }\) for all \(z\ge T_{\psi _c}\). Finally, by applying Theorem 18 we conclude (24) with \(C:=\max _{z\in {\mathbb {R}}}\eta _{b}(z)/\eta _{t_0+ch}(z)\) where \(t_0:=\max \{T_v,T_{\psi _c} \}.\) \(\square \)