1 Introduction

This paper investigates reaction diffusion equations of the type

$$\begin{aligned} u_{t}=\varDelta u+f(u),\quad (t,x)\in \mathbb {R}\times \mathbb {R}^{N}, \end{aligned}$$
(1)

where \(N\in \mathbb {N}\), \(u_{t}=\frac{\partial u}{\partial t}\) and \(\varDelta \) denotes the Laplace operator with respect to the space variables \(x\in \mathbb {R}^{N}\). The nonlinear reaction term f(u) is of the “ignition temperature” type, that is, \(f:[0,1]\rightarrow \mathbb {R}\) is a \(C^{1}\) function such that

$$\begin{aligned} \exists ~\theta \in (0,1),~f\equiv 0~\mathrm{on}~[0,\theta ]\cup \{1\},~f>0~\mathrm{on}~(\theta ,1)~\mathrm{and}~f'(1)<0. \end{aligned}$$
(2)

Such a profile can be derived from the Arrhenius kinetis with a cut-off for low temperatures and from the law of mass action. The real number \(\theta \) is the ignition temperature, below which no reaction happens.

In any dimension \(N\ge 1\), standard planar traveling fronts are solutions of the type

$$\begin{aligned} u(t,x)=\phi _{f}\left( x\cdot e-c_{f}t\right) , \end{aligned}$$

where e is any given unit vector of \(\mathbb {R}^{N}\), \(c_{f}\in \mathbb {R}\) is the propagation speed and \(\phi _{f}:\mathbb {R}\rightarrow [0,1]\) is the propagation profile, such that

$$\begin{aligned} \begin{aligned}&\phi ^{\prime \prime }_{f}+c_{f}\phi ^{\prime }_{f}+f\left( \phi _{f}\right) =0\quad \mathrm{in}~\mathbb {R},\\&\phi _{f}(-\infty )=1~~\mathrm{and}~~\phi _{f}(+\infty )=0. \end{aligned} \end{aligned}$$
(3)

The profile \(\phi _{f}\) is then a heteroclinic connection between the state 0 and the stable state 1. The level sets of such traveling fronts are parallel hyperplanes which are orthogonal to the direction of the propagation e. These fronts are invariant in the moving frame with speed \(c_{f}\) in the direction e. It is well known [1] that such front exists and is unique up to translation. Besides, the speed \(c_{f}\) is positive which has the sign of \(\int _{0}^{1}f(s)ds\) [5] and the function \(\phi _{f}\) is decreasing.

In \(\mathbb {R}^{N}\) with \(N\ge 2\), propagating wave fronts contains more types of traveling fronts except planar traveling fronts, such as V-shaped traveling fronts in two-dimensional spaces, pyramidal traveling fronts with non-axisymmetric shape in three-dimensional spaces and conical-shaped axisymmetric traveling fronts in high-dimensional spaces. The profiles of these fronts are still invariant in a moving frame with constant speed. But they have non-planar level sets. For instance, (1) admits the conical-shaped fronts of the type

$$\begin{aligned}u(t,x)=\phi (|x'|,x_{N}-c t),\end{aligned}$$

where \(x'=(x_1,\ldots ,x_{N-1})\) and \(|x'|=(x_1^2+\cdots +x_{n-1}^2)^{1/2}\) whose profiles are invariant and which have non-planar level sets. For the existence, uniqueness, stability and other qualitative properties of these non-planar traveling fronts, we refer to [7, 8, 12,13,14, 24, 25, 33,34,35,36] and the references therein.

As we introduced above, Eq. (1) admits many types of traveling fronts. However, they have some common properties. For instance, the solutions u converge to the equilibrium states 0 or 1 far away from their moving or stationary level sets, uniformly in time. Their common properties led us to ask whether it is possible to introduce a more general notion of traveling fronts to include all types of waves. Berestycki and Hamel [3, 4] give an affirmative answer. They introduce the general notion of transition fronts. Before we describe the definition of transition fronts, we firstly introduce some notions. For any two subsets A and B of \(\mathbb {R}^{N}\) and for \(x\in \mathbb {R}^{N}\), we set

$$\begin{aligned} d(A,B)=\inf \{\vert x-y\vert , (x,y)\in A\times B\} \end{aligned}$$
(4)

and \(d(x,A)=d(\{x\},A)\), where \(\vert \cdot \vert \) is the Euclidean norm in \(\mathbb {R}^{N}\). Let \((\varOmega _{t}^{-})_{t\in \mathbb {R}}\) and \((\varOmega _{t}^{+})_{t\in \mathbb {R}}\) be two families of open nonempty subsets of \(\mathbb {R}^{N}\), which satisfy

$$\begin{aligned} \forall ~t\in \mathbb {R},\quad {\left\{ \begin{array}{ll} \varOmega _{t}^{-}\bigcap \varOmega _{t}^{+}=\emptyset ,\\ \partial \varOmega _{t}^{-}=\partial \varOmega _{t}^{+}=:\varGamma _{t},\\ \varOmega _{t}^{-}\bigcup \varGamma _{t}\bigcup \varOmega _{t}^{+}=\mathbb {R}^{N},\\ \sup \{d(x,\varGamma _{t})\big \vert x\in \varOmega _{t}^{+}\}= \sup \{d(x,\varGamma _{t})\big \vert x\in \varOmega _{t}^{-}\}=+\infty \end{array}\right. } \end{aligned}$$
(5)

and

$$\begin{aligned} \begin{aligned}&\inf \{\sup \{d(y,\varGamma _{t});y\in \varOmega _{t}^{+},\vert y-x\vert \le r\}\big \vert t\in \mathbb {R},x\in \varGamma _{t}\}=+\infty ,\\&\inf \{\sup \{d(y,\varGamma _{t});y\in \varOmega _{t}^{-},\vert y-x\vert \le r\}\big \vert t\in \mathbb {R},x\in \varGamma _{t}\}=+\infty , \end{aligned} \quad r\rightarrow +\infty . \end{aligned}$$
(6)

Notice that the condition (5) implies that the interface \(\varGamma _{t}\) is not empty for every \(t\in \mathbb {R}\).

Definition 1

(See [3, 4]) For problem (1), a transition front connecting 0 and 1 is a classical solution \(u:\mathbb {R}\times \mathbb {R}^{N}\rightarrow (0,1)\) for which there exist some sets \((\varOmega _{t}^{\pm })_{t\in \mathbb {R}}\) and \((\varGamma _{t})_{t\in \mathbb {R}}\) satisfying (5) and (6), and, for every \(\varepsilon >0\), there exists \(M\ge 0\) such that

$$\begin{aligned} \begin{aligned}&\forall ~t\in \mathbb {R},~\forall ~x\in \varOmega _{t}^{+},\quad d(x,\varGamma _{t})\ge M\Rightarrow u(t,x)\ge 1-\varepsilon ,\\&\forall ~t\in \mathbb {R},~\forall ~x\in \varOmega _{t}^{-},\quad d(x,\varGamma _{t})\ge M\Rightarrow u(t,x)\le \varepsilon . \end{aligned} \end{aligned}$$
(7)

Furthermore, u is said to have a global mean speed \(\varLambda ~(\ge 0)\) if

$$\begin{aligned} \frac{d(\varGamma _{t},\varGamma _{s})}{\vert t-s\vert }\rightarrow \varLambda \quad as~\vert t-s\vert \rightarrow +\infty . \end{aligned}$$
(8)

Remark 1

Notice that, for a given transition front u connecting 0 and 1, the sets \((\varOmega _{t}^{\pm })_{t\in \mathbb {R}}\) and \((\varGamma _{t})_{t\in \mathbb {R}}\) are not uniquely determined. In fact, for any sets \((\widetilde{\varGamma }_{t})_{t\in \mathbb {R}}\), if

$$\begin{aligned} \sup _{t\in \mathbb {R}}\max \left( \sup _{x\in \varGamma _t} d(x,\widetilde{\varGamma }_{t}),\ \sup _{x\in \widetilde{\varGamma }_t}d(x,\varGamma _t)\right) <+\infty , \end{aligned}$$

then the family \((\widetilde{\varGamma }_{t})_{t\in \mathbb {R}}\) with corresponding sets \((\widetilde{\varOmega }_{t}^{\pm })_{t\in \mathbb {R}}\) also satisfies (5), (6) and (7). That is, the solution u is also a transition front connecting 0 and 1 with the families \((\widetilde{\varOmega }_{t}^{\pm })_{t\in \mathbb {R}}\) and \((\widetilde{\varGamma }_{t})_{t\in \mathbb {R}}\).

Notice furthermore that for any transition front u connecting 0 and 1, the interfaces \((\varGamma _{t})_{t\in \mathbb {R}}\) have uniformly bounded local oscillations, that is

$$\begin{aligned} \forall ~\sigma >0,~\sup \left\{ d(\varGamma _{t},\varGamma _{s}),~t,s\in \mathbb {R},~\vert t-s\vert \le \sigma \right\} <+\infty . \end{aligned}$$
(9)

In fact, it is shown in Lemma 3 and Remark 3 of [10], in the case of reaction–diffusion equations (1) with nonlinearity f satisfying \(f(u)>0\) for \(u\in (1-\delta ,1)\), where \(0<\delta <1\). Obviously, the assumptions of nonlinear reaction term f in this paper (see (2)) satisfy the above condition with \(\delta =1-\theta \).

In [3, 4, 11], the authors have showed that all the known standard traveling fronts (planar and non-planar traveling fronts) are transition fronts in the sense of Definition 1. In particular, Hamel [11] proved that for Eq. (1) with bistable nonlinearity there exist new types of transition fronts in \(\mathbb {R}^{N}\) which are not invariant in any frame as time runs. This property is different from standard traveling fronts which are invariant in a moving frame with constant speed. It also shows the broadness of Definition 1. In recent years, many papers have been devoted to the investigation of the existence and stability of transition fronts. For bistable transition fronts, we refer to [3, 4, 10, 11]. For Fisher-KPP transition fronts, the readers can see [15, 16, 21,22,23, 28, 31, 38]. Transition fronts for equations with combustion nonlinearity, the investigations mainly focus on the case of the heterogeneous equations in \(\mathbb {R}\), see [19, 20, 29, 30, 32, 37, 39, 40]. In this paper, we prove that even the homogeneous combustion equation (1) in \(\mathbb {R}^{N}\)\((N\ge 1)\) also has many deep properties, such as the existence of new transition fronts and general estimates shared by all transition fronts.

The first main result of this paper proves the existence and uniqueness of the global mean speed for any transition fronts connecting the state 0 and the stable state 1, regardless of the shape of the level sets of the transition fronts.

Theorem 1

For problem (1), any transition front u connecting 0 and 1 has a global mean speed \(\varLambda \). Furthermore, this global mean speed \(\varLambda \) is equal to \(c_{f}\).

The second result of this paper gives a characterization of the planar fronts \(\phi _{f}(x\cdot e-c_{f}t)\) among the more general class of almost-planar transition fronts introduced in [4], and defined as follows.

Definition 2

(See [4, 11]) A transition front u in the sense of Definition 1 is called almost-planar if, for every \(t\in \mathbb {R}\), the set \(\varGamma _{t}\) can be chosen as the hyperplane

$$\begin{aligned} \varGamma _{t}=\left\{ x\in \mathbb {R}^{N}\big \vert x\cdot e_{t}=\xi _{t}\right\} \end{aligned}$$

for some vector \(e_{t}\) of the unit sphere \(\mathbb {S}^{N-1}\) and some real number \(\xi _{t}\).

From the definition, we can easily see that the level sets of almost-planar fronts are in some sense close to hyperplanes, even if they are not a priori assumed to be planar. The following theorem shows that planar fronts \(\phi _{f}(x\cdot e-c_{f}t)\) for problem (1) fall within the more general class of almost-planar fronts.

Theorem 2

For problem (1), any almost-planar transition front u connecting 0 and 1 is planar, that is, there exist a unit vector e of \(\mathbb {R}^{N}\) and a real number \(\xi \) such that

$$\begin{aligned} u(t,x)=\phi _{f}(x\cdot e-c_{f}t+\xi )\quad for~all~(t,x)\in \mathbb {R}\times \mathbb {R}^{N}. \end{aligned}$$

Thirdly, we show the broadness of transition fronts. In other words, we prove the existence of new types transition fronts of the Eq. (1), which are not invariant as time runs in any moving frame. Recall that the profiles of standard traveling fronts are invariant in a moving frame with constant speed.

Theorem 3

Let \(N\ge 2\). The problem (1) admits transition fronts u connecting 0 and 1 which satisfy the following property: there is no function \(\varPhi :\mathbb {R}^{N}\rightarrow (0,1)\) (independent of t) for which there would be some families \((R_{t})_{t\in \mathbb {R}}\) and \((x_{t})_{t\in \mathbb {R}}\) of rotations and points in \(\mathbb {R}^{N}\) such that \(u(t,x)=\varPhi (R_{t}(x-x_{t}))\) for all \((t,x)\in \mathbb {R}\times \mathbb {R}^{N}\).

Finally, we establish the time monotonicity of the transition front u.

Theorem 4

For problem (1), any transition front u connecting 0 and 1 is monotone increasing in time t. That is, \(u_{t}>0\) for all \((t,x)\in \mathbb {R}\times \mathbb {R}^{N}\).

In fact, in order to prove Theorem 4, it is sufficient to prove that the transition front u is an invasion of the state 0 by the state 1, in the sense that the sets \((\varOmega ^{\pm }_t)_{t\in \mathbb {R}}\) can be chosen so that

$$\begin{aligned} \varOmega _{t}^{+}\subset \varOmega _{s}^{+}~\mathrm{for~all}~t<s~\mathrm{and}~d(\varGamma _{t},\varGamma _{s})\rightarrow +\infty ~\mathrm{as}~\vert t-s\vert \rightarrow +\infty , \end{aligned}$$

since it is easy to check that the problem (1) and the nonlinearity f satisfy all assumptions of [4, Theorem 1.11]. Similar to [10], it follows from Theorem 1 and Lemma 1 (see Sect. 2) that u is an invasion in the above sense with the families \(\left( \widehat{\varOmega }_{t}^{\pm }\right) _{t\in \mathbb {R}}\) and \(\left( \widehat{\varGamma }_{t}\right) _{t\in \mathbb {R}}\), where for some constant \(\tau _{0}>0\),

$$\begin{aligned} \begin{array}{ll} \widehat{\varOmega }_{k\tau _{0}+t}^{\pm }=\varOmega _{k\tau _{0}}^{\pm }&{}\quad \mathrm{for~any~}k\in \mathbb {Z}~\mathrm{and}~0\le t<\tau _{0},\\ \widehat{\varGamma }_{t}:=\partial \widehat{\varOmega }_{t}^{+}=\widehat{\varOmega }_{t}^{-}&{}\quad \mathrm{for~any~}t\in \mathbb {R}. \end{array} \end{aligned}$$

Now we give a brief stated on the methods of our proofs. Firstly, in order to prove the existence and the uniqueness of the global mean speed of the transition fronts connecting 0 and 1, we need introduce two radially symmetric functions and show their dynamical properties, see Lemmas 1 and 2 below. Secondly, using the one-dimensional stability of the planar front and parabolic Liouville type result of Berestycki and Hamel [3, Theorem 3.1], we show that the planar fronts can be characterized by the more general class of almost-planar transition fronts. Thirdly, by mixing three planar fronts moving in three different directions, we show that the new transition fronts exist in dimension \(N=2\). And by trivially extending the two-dimensional solutions in the variables \(x_{3},\ldots ,x_{N}\), we obtain that the new transition fronts exist in all dimensions \(N\ge 3\).

Here we would like to point out that the main results of this paper (Theorems 1, 2, 3 and 4) are similar to those established for Eq. (1) with bistable nonlinearity by Hamel [11] and Guo and Hamel [10], where the reaction term \(f:[0,1]\rightarrow \mathbb {R}\) is a \( C^1\) function such that

$$\begin{aligned} f(0)=f(1)=0,\ f^\prime (0)<0\ \mathrm{and}\ f^\prime (1)<0. \end{aligned}$$

But in this paper we treat the combustion case, in particular, the reaction term f satisfies \(f(u)=0\) for any \(u\in [0,\theta ]\) with some \(\theta \in (0,1)\), which is essentially different from the assumption \(f^\prime (0)<0\) in the bistable case. Since the signs of \(f'(0)\) and \(f'(1)\) play important roles in the estimates of speeds and constructing the super-sub solutions, some new difficulties occur in the combustion reaction diffusion equations. To overcome these difficulties, we need some new techniques and establish some new estimates. See Lemmas 1, 2, 4 and Proposition 1 below for more details.

The rest of this paper is organized as follows. Section 2 proves the existence and the uniqueness of the global mean speed among all transition fronts. That is, we give the proof of Theorem 1. In Sect. 3, we prove Theorem 2. That is, we give a characterization of the planar fronts among the more general class of almost-planar transition fronts. In Sect. 4, we construct new types transition fronts. That is, we are devoted to the proof of Theorem 3.

2 The Global Mean Speed

In this section, we prove that any transition front of the Eq. (1) has a global mean speed and this speed is unique. We first introduce auxiliary notations for some radially symmetric functions and we show some of their dynamical properties. The following two key properties, Lemmas 1 and 2 below, will provide a sharp lower bound and a upper bound for the speed of the interfaces \(\varGamma _{t}\) of any transition front connecting 0 and 1 for the problem (1), respectively.

In the following, let \(\theta<\beta <1\). For any \(R>0\), let \(v_{R}^{f}\) denote the solution of the Cauchy problem

$$\begin{aligned} \big (v_{R}^{f}\big )_{t}=\varDelta v_{R}^{f}+f\big (v_{R}^{f}\big ),\quad t>0,\quad x\in \mathbb {R}^{N}, \end{aligned}$$
(10)

with initial value

$$\begin{aligned} v_{R}^{f}(0,x)= {\left\{ \begin{array}{ll} \beta ,~~~\vert x\vert <R,\\ 0,~~~\vert x\vert \ge R. \end{array}\right. } \end{aligned}$$
(11)

Lemma 1

There is \(R>0\) such that the following holds: for any \(\varepsilon \in (0,c_{f}]\), there is \(T_{\varepsilon }>0\) such that

$$\begin{aligned} v_{R}^{f}(t,x)\ge \beta \quad for~all~t\ge T_{\varepsilon }~and~\vert x\vert \le (c_{f}-\varepsilon )t. \end{aligned}$$
(12)

In fact,

$$\begin{aligned} v_{R}^{f}(t,\cdot )\rightarrow 1\quad uniformly~in~\left\{ x\in \mathbb {R}^{N}\big \vert \vert x\vert \le (c_{f}-\varepsilon )t\right\} ~as~t\rightarrow +\infty . \end{aligned}$$
(13)

Proof

Let g be any given \(C^{1}([0,1])\) function which satisfies

$$\begin{aligned} \begin{aligned}&g(0)=g(\theta )=g(1)=0,~g^{\prime }(0)<0,~g^{\prime }(1)<0,~g^{\prime }(\theta )>0,\\&g<0~\mathrm{on}~(0,\theta ),~0<g\le f~\mathrm{on}~(\theta ,1), \int _{0}^{1}g(s)ds>0 \end{aligned} \end{aligned}$$

and

$$\begin{aligned} 0<c_{f}-c_{g}\le \frac{\varepsilon }{2}, \end{aligned}$$
(14)

where \(c_{g}\) is the wave speed of the planar front \(\phi _{g}\) which satisfies (3) with the nonlinearity g. In fact, g is of the bistable type. Such fronts exist, see [2, 9, 17]. It is easy to see that \(f\ge g\) on [0, 1]. Then the comparison principle implies that

$$\begin{aligned} 1\ge v_{R}^{f}(t,x)\ge v_{R}^{g}(t,x),\quad \forall ~(t,x)\in [0,+\infty )\times \mathbb {R}^{N}. \end{aligned}$$
(15)

For the solution \(v_{R}^{g}\) of the equation (10)-(11) with replacing f by g, it follows from Lemma 4.1 of [11] that we have

$$\begin{aligned} v_{R}^{g}(t,\cdot )\rightarrow 1\quad \mathrm{uniformly~in}~\left\{ x\in \mathbb {R}^{N}\big \vert \vert x\vert \le \left( c_{g}-\frac{\varepsilon }{2}\right) t\right\} ~\mathrm{as}~t\rightarrow +\infty . \end{aligned}$$

Inequalities (14) and (15), together with the above formula, yield that (13) holds. This completes the proof. \(\square \)

Lemma 2

For any \(\varepsilon >0\), there exist some positive real numbers \(\alpha _{\varepsilon }\), \(T_{\varepsilon }\) and \(R_{\varepsilon }\) such that for all \(R\ge R_{\varepsilon }\), the solution \(w_{R}\) of the following Cauchy problem

$$\begin{aligned} (w_{R})_{t}=\varDelta w_{R}+f(w_{R}),\quad t>0,\quad x\in \mathbb {R}^{N}, \end{aligned}$$

with initial value

$$\begin{aligned} w_{R}(0,x)= {\left\{ \begin{array}{ll} \alpha _{\varepsilon },~~\vert x\vert <R,\\ 1,~~~~\vert x\vert \ge R \end{array}\right. } \end{aligned}$$

satisfies

$$\begin{aligned} w_{R}(t,x)\le 3\alpha _{\varepsilon }\quad for~all~T_{\varepsilon }\le t\le \frac{R}{c_{f}+\varepsilon }~and~\vert x\vert \le R-(c_{f}+\varepsilon )t. \end{aligned}$$
(16)

Proof

Let \(\delta \) be chosen so that

$$\begin{aligned} 0<\delta <\frac{\theta }{2}~~\mathrm{and}~~f'\le \frac{f'(1)}{2}~~\mathrm{on}~[1-\delta ,1]. \end{aligned}$$
(17)

Since \(\phi ''_{f}(s)\sim \nu e^{-c_{f}s}\) as \(s\rightarrow +\infty \) with \(\nu >0\), one can choose \(C>0\) such that

$$\begin{aligned} \phi _{f}\ge 1-\delta ~\mathrm{on}~(-\infty ,-C],~\phi _{f}\le \delta ~\mathrm{on}~[C,+\infty )~\mathrm{and}~\phi ''_{f}\ge 0~\mathrm{on}~[C,+\infty ). \end{aligned}$$
(18)

Since \(\phi '_{f}\) is negative and continuous on \(\mathbb {R}\), there is \(\kappa >0\) such that

$$\begin{aligned} -\phi '_{f}\ge \kappa >0\quad \mathrm{on}~[-C,C]. \end{aligned}$$
(19)

Set \(L=\max \limits _{u\in [0,1]}\vert f'(u)\vert \). For every \(\varepsilon >0\), let

$$\begin{aligned} 0<\alpha _{\varepsilon }<\min \left( \frac{\theta }{4},\frac{\kappa \varepsilon }{8L}\right) . \end{aligned}$$

Choose \(D_{\varepsilon }>0\) such that

$$\begin{aligned} \phi _{f}\ge 1-2\alpha _{\varepsilon }~\mathrm{on}~(-\infty ,-D_{\varepsilon }]~~\mathrm{and}~~\phi _{f}\le \alpha _{\varepsilon }~\mathrm{on}~[D_{\varepsilon },+\infty ). \end{aligned}$$
(20)

Let \(\rho _{\alpha _{\varepsilon }}\) be the solution of the following ordinary differential equation

$$\begin{aligned} \begin{aligned}&\rho ^{\prime }_{\alpha _{\varepsilon }}(t)=f(\rho _{\alpha _{\varepsilon }}),\\&\rho _{\alpha _{\varepsilon }}(0)=\alpha _{\varepsilon }. \end{aligned} \end{aligned}$$

Since \(\alpha _{\varepsilon }\in (0,\theta )\), f is Lipschitz-continuous and \(f\equiv 0\) on \([0,\theta ]\), then \(\rho _{\alpha _{\varepsilon }}(t)\equiv \alpha _{\varepsilon }\) by the existence and uniqueness of solution of the ordinary differential equation. It follows from the maximum principle and (2) that for any \(R>0\),

$$\begin{aligned} 0\le \rho _{\alpha _{\varepsilon }}(t)\le w_{R}(t,x)\le 1\quad \mathrm{for~all}~t\ge 0,\quad x\in \mathbb {R}^{N}. \end{aligned}$$

Then the following inequality holds

$$\begin{aligned} (w_{R}-\rho _{\alpha _{\varepsilon }})_{t}\le \varDelta (w_{R}-\rho _{\alpha _{\varepsilon }})+L(w_{R}-\rho _{\alpha _{\varepsilon }}). \end{aligned}$$

Thus for the above equation, the assumptions of the initial value yield

$$\begin{aligned} 0\le w_{R}(t,x)-\rho _{\alpha _{\varepsilon }}(t)\le \frac{e^{Lt}}{(4\pi t)^{\frac{N}{2}}}\int \limits _{\vert y\vert \ge R}e^{-\frac{\vert x-y\vert ^{2}}{4t}}dy\quad \mathrm{for~all}\quad t>0~\mathrm{and}~x\in \mathbb {R}^{N}. \end{aligned}$$

Therefore, if \(0<B\le R\) and \(\vert x\vert \le R-B\), one infers that

$$\begin{aligned} 0\le w_{R}(t,x)-\rho _{\alpha _{\varepsilon }}(t)\le \frac{e^{Lt}}{(4\pi )^{\frac{N}{2}}}\int \limits _{\vert z\vert \ge \frac{B}{\sqrt{t}}}e^{-\vert z\vert ^{2}}dz. \end{aligned}$$

Thus, take a real number \(T>0\) and there exists \(B>0\) such that for all \(R\ge B\) and \(\vert x\vert \le R-B\),

$$\begin{aligned} w_{R}(T,x)-\rho _{\alpha _{\varepsilon }}(T)\le \alpha _{\varepsilon }, \end{aligned}$$

whence

$$\begin{aligned} w_{R}(T,x)\le \rho _{\alpha _{\varepsilon }}(T)+\alpha _{\varepsilon }=2\alpha _{\varepsilon }\quad \mathrm{for~all}~R\ge B~\mathrm{and}~\vert x\vert \le R-B. \end{aligned}$$
(21)

It is elementary to check that for every \(\varepsilon >0\), there is a \(C^{2}\) function \(h_{\varepsilon }:[0,+\infty )\rightarrow \mathbb {R}\) satisfying the following properties:

$$\begin{aligned} \begin{aligned}&0\le h^{\prime }_{\varepsilon }\le 1\quad \mathrm{on}~[0,+\infty ),\\&h^{\prime }_{\varepsilon }=0\quad \mathrm{on~a~neighborhood~of}~0,\\&h_{\varepsilon }(r)=r\quad \mathrm{on}~[H_{\varepsilon },+\infty )~\mathrm{for~some}~H_{\varepsilon }>0,\\&\frac{(N-1)h^{\prime }_{\varepsilon }(r)}{r}+h^{\prime \prime }_{\varepsilon }(r)\le \frac{\varepsilon }{4}\quad \mathrm{on}~[0,+\infty ). \end{aligned} \end{aligned}$$
(22)

Notice in particular that

$$\begin{aligned} r\le h_{\varepsilon }(r)\le r+h_{\varepsilon }(0)\quad \mathrm{for~all}~r\ge 0. \end{aligned}$$
(23)

We choose \(T_{\varepsilon }>T>0\) such that

$$\begin{aligned} \frac{\varepsilon t}{2}\ge h_{\varepsilon }(0)+B+2D_{\varepsilon }\quad \mathrm{for~all}~t\ge T_{\varepsilon }, \end{aligned}$$
(24)

and \(R_{\varepsilon }>0\) such that

$$\begin{aligned} R_{\varepsilon }\ge \max (B,(c_{f}+\varepsilon )T_{\varepsilon })\quad \mathrm{and}\quad \frac{\varepsilon R_{\varepsilon }}{2(c_{f}+\varepsilon )}\ge B+D_{\varepsilon }+C+H_{\varepsilon }. \end{aligned}$$
(25)

In the sequel, R is arbitrary real number such that

$$\begin{aligned} R\ge R_{\varepsilon }. \end{aligned}$$
(26)

For all \((t,x)\in \mathbb {R}\times \mathbb {R}^{N}\), we set

$$\begin{aligned} \overline{W}(t,x)=\min \left( \phi _{f}(\bar{\xi }(t,x))+2\alpha _{\varepsilon },1\right) , \end{aligned}$$

where

$$\begin{aligned} \bar{\xi }(t,x)=-h_{\varepsilon }(\vert x\vert )-\left( c_{f}+\frac{\varepsilon }{2}\right) (t-T)+R-B-D_{\varepsilon }. \end{aligned}$$

In the set \(\varSigma =\left[ T,\frac{R}{c_{f}+\varepsilon }\right] \times \mathbb {R}^{N}\), let us then check that \(\overline{W}\) is a supersolution for the problem satisfied by \(w_{R}\).

Since \(f(1)=0\), it is sufficient to check that

$$\begin{aligned} \mathscr {L}(t,x)=\overline{W}_{t}(t,x)-\varDelta \overline{W}(t,x)-f(\overline{W}(t,x))\ge 0~~\mathrm{for~all}~(t,x)\in \varSigma ~\mathrm{such~that}~\overline{W}(t,x)<1. \end{aligned}$$

Since \(\phi _{f}\) is of class \(C^{2}\) and h vanishes in the neighborhood of 0, then \(\overline{W}(t,x)=\phi _{f}(\bar{\xi }(t,x))+2\alpha _{\varepsilon }\) is of class \(C^{2}\) in the set where \(\overline{W}(t,x)<1\).

In this paragraph, let (tx) be any point in \(\varSigma \) such that \(\overline{W}(t,x)<1\). Since \(\phi ^{\prime \prime }_{f}+c_{f}\phi '_{f}+f(\phi _{f})=0\) in \(\mathbb {R}\), then by (22) and \(\phi '_{f}\le 0\), there holds

$$\begin{aligned} \mathscr {L}(t,x)=&f(\phi _{f}(\bar{\xi }(t,x)))-f(\overline{W}(t,x))+(1-(h^{\prime }_{\varepsilon }(\vert x\vert ))^{2})\phi ^{\prime \prime }_{f}(\bar{\xi }(t,x))\\&-\left( \frac{\varepsilon }{2}-\frac{(N-1)h^{\prime }_{\varepsilon }(\vert x\vert )}{\vert x\vert }-h^{\prime \prime }_{\varepsilon }(\vert x\vert )\right) \phi ^{\prime }_{f}(\bar{\xi }(t,x))\\ \ge&f(\phi _{f}(\bar{\xi }(t,x)))-f(\overline{W}(t,x)) -\frac{\varepsilon }{4}\phi ^{\prime }_{f}(\bar{\xi }(t,x)) +(1-(h^{\prime }_{\varepsilon }(\vert x\vert ))^{2})\phi ^{\prime \prime }_{f}(\bar{\xi }(t,x)). \end{aligned}$$

Firstly, if \(\bar{\xi }(t,x)\le -C\), then (18) and the definition of \(\overline{W}\) yield \(1-\delta \le \phi _{f}(\bar{\xi }(t,x))\le \overline{W}(t,x)<1\). Whence by (17), one gets

$$\begin{aligned} f(\phi _{f}(\bar{\xi }(t,x)))-f(\overline{W}(t,x))\ge -f'(1)\alpha _{\varepsilon }. \end{aligned}$$

In addition, it follows from (25) and (26) that the inequalitues \(\bar{\xi }(t,x)\le -C\) and \(T\le t\le \frac{R}{c_{f}+\varepsilon }\) yield

$$\begin{aligned} h_{\varepsilon }(\vert x\vert )\ge -\left( c_{f}+\frac{\varepsilon }{2}\right) (t-T)+R-B-D_{\varepsilon }+C \ge \frac{\varepsilon R}{2(c_{f}+\varepsilon )}-B-D_{\varepsilon }+C\ge H_{\varepsilon }. \end{aligned}$$

From the properties (22), the inequality \(h_{\varepsilon }(\vert x\vert )\ge H_{\varepsilon }\) implies that \(h'_{\varepsilon }(\vert x\vert )= 1\). Therefore, if \(\bar{\xi }(t,x)\le -C\), then \(\phi '_{f}\le 0\) implies

$$\begin{aligned} \mathscr {L}(t,x)\ge -f'(1)\alpha _{\varepsilon } -\frac{\varepsilon }{4}\phi '_{f}(\bar{\xi }(t,x))\ge 0. \end{aligned}$$

Secondly, if \(\bar{\xi }(t,x)\ge C\), then by (18), \(\phi _{f}(\bar{\xi }(t,x))\le \delta \). Thus,

$$\begin{aligned} 0<\phi _{f}(\bar{\xi }(t,x))\le \overline{W}(t,x)\le \delta +2\alpha _{\varepsilon }<\theta . \end{aligned}$$

Since \(f=0\) on \([0,\theta ]\), \(\phi ''_{f}\ge 0\) on \([C,+\infty )\) from (18), \(0\le h'_{\varepsilon }(\vert x\vert )\le 1\) on \([0,+\infty )\) and \(\phi '_{f}\le 0\) on \(\mathbb {R}\), one gets that, if \(\bar{\xi }(t,x)\ge C\), then

$$\begin{aligned} \mathscr {L}(t,x)\ge -\frac{\varepsilon }{4}\phi '_{f}(\bar{\xi }(t,x))+(1-(h'_{\varepsilon }(\vert x\vert ))^{2})\phi ''_{f}(\bar{\xi }(t,x))\ge 0. \end{aligned}$$

Lastly, if \(-C\le \bar{\xi }(t,x)\le C\), then

$$\begin{aligned} f(\phi _{f}(\bar{\xi }(t,x)))-f(\overline{W}(t,x))\ge -2 L\alpha _{\varepsilon }, \end{aligned}$$

recall that \(L=\max \limits _{u\in [0,1]}\vert f'(u)\vert \). It follows from (24) and (26) that \(\bar{\xi }(t,x)\le C\) and \(T\le t\le \frac{R}{c_{f}+\varepsilon }\) imply

$$\begin{aligned} h_{\varepsilon }(\vert x\vert )\ge -\left( c_{f}+\frac{\varepsilon }{2}\right) (t-T)+R-B-D_{\varepsilon }-C \ge \frac{\varepsilon R}{2(c_{f}+\varepsilon )}-B-D_{\varepsilon }-C\ge H_{\varepsilon }. \end{aligned}$$

Thus by (22), \(h'_{\varepsilon }(\vert x\vert )=1\). Consequently, it follows from the definition of \(\alpha _{\varepsilon }\) and (19) that

$$\begin{aligned} \mathscr {L}(t,x)\ge -2L\alpha _{\varepsilon }+\frac{\kappa \varepsilon }{4}\ge 0. \end{aligned}$$

On the other hand, at the time T, it follows from (21), (24), (26) and the definition of \(\overline{W}\) that

$$\begin{aligned} w_{R}(T,x)\le 2\alpha _{\varepsilon }\le \overline{W}(T,x)\quad \mathrm{for~all}~\vert x\vert \le R-B. \end{aligned}$$

If \(\vert x\vert \ge R-B\), then \(h_{\varepsilon }(\vert x\vert )\ge \vert x\vert \ge R-B\) from (23), whence \(\bar{\xi }(T,x)\le -D_{\varepsilon }\) and

$$\begin{aligned} \overline{W}(T,x)=\min \left( \phi _{f}(\bar{\xi }(T,x))+2\alpha _{\varepsilon },1\right) \ge \min \left( (1-2\alpha _{\varepsilon })+2\alpha _{\varepsilon },1\right) =1\ge w_{R}(T,x) \end{aligned}$$

from (20) and the fact that \(w_{R}\le 1\) on \((0,+\infty )\times \mathbb {R}^{N}\). Thus

$$\begin{aligned} w_{R}(T,x)\le \overline{W}(T,x)\quad \mathrm{for~all}~x\in \mathbb {R}^{N}. \end{aligned}$$

As a conclusion, the maximum principle implies that, for all \(T\le t\le \frac{R}{c_{f}+\varepsilon }\) and \(x\in \mathbb {R}^{N}\),

$$\begin{aligned} w_{R}(t,x)\le \overline{W}(t,x)\le \phi _{f}(\bar{\xi }(t,x))+2\alpha _{\varepsilon }. \end{aligned}$$

For all \(T_{\varepsilon }\le t\le \frac{R}{c_{f}+\varepsilon }\) and \(\vert x\vert \le R-(c_{f}+\varepsilon )t\), there hold

$$\begin{aligned} h_{\varepsilon }(\vert x\vert )\le \vert x\vert +h_{\varepsilon }(0)\le R-(c_{f}+\varepsilon )t+h_{\varepsilon }(0) \end{aligned}$$

and

$$\begin{aligned} \bar{\xi }(t,x)\ge&-R+(c_{f}+\varepsilon )t-h_{\varepsilon }(0) -\left( c_{f}+\frac{\varepsilon }{2}\right) (t-T) +R-B-D_{\varepsilon }\\ \ge&\frac{\varepsilon t}{2}-h_{\varepsilon }(0)-B-D_{\varepsilon }\\ \ge&D_{\varepsilon } \end{aligned}$$

from (24). Thus, (20) yields \(\phi _{f}(\bar{\xi }(t,x))\le \alpha _{\varepsilon }\). Whence, if \(T_{\varepsilon }\le t\le \frac{R}{c_{f}+\varepsilon }\) and \(\vert x\vert \le R-(c_{f}+\varepsilon )t\), then

$$\begin{aligned} w_{R}(t,x)\le \phi _{f}(\bar{\xi }(t,x))+2\alpha _{\varepsilon }\le \alpha _{\varepsilon }+2\alpha _{\varepsilon }=3\alpha _{\varepsilon }. \end{aligned}$$

This completes the proof. \(\square \)

Proof of Theorem 1

Let u be any transition front of problem (1) which connects the equilibrium points 0 and 1. For any \(\varepsilon \in (0,c_{f}]\), let \(\alpha _{\varepsilon }\) be defined as in Lemma 2 and \(\theta<\beta <1\). It follows from Definition 1 that there is \(M\ge 0\) such that

$$\begin{aligned} \begin{aligned}&\forall ~t\in \mathbb {R},~\forall ~x\in \overline{\varOmega _{t}^{+}},\quad d(x,\varGamma _{t})\ge M\Rightarrow \beta \le u(t,x)<1,\\&\forall ~t\in \mathbb {R},~\forall ~x\in \overline{\varOmega _{t}^{-}},\quad d(x,\varGamma _{t})\ge M\Rightarrow 0<u(t,x)\le \alpha _{\varepsilon }. \end{aligned} \end{aligned}$$
(27)

Let \(R>0\) be as in Lemma 1. Without loss of generality, one can assume that \(R\ge M\) (since the functions \(v_{R}^{f}\) are nondecreasing with respect to the parameter \(R>0\)). By (6), there exists a real number \(r>0\) such that

$$\begin{aligned} \forall ~t\in \mathbb {R},\quad \forall ~x\in \varGamma _{t},\quad \exists ~y^{\pm }\in \varOmega _{t}^{\pm },\quad \vert x-y^{\pm }\vert \le r\quad \mathrm{and}\quad d(y^{\pm },\varGamma _{t})\ge 2R. \end{aligned}$$
(28)

Our goal is to prove

$$\begin{aligned} \frac{d(\varGamma _{t},\varGamma _{s})}{\vert t-s\vert }\rightarrow c_{f}\quad \mathrm{as}~\vert t-s\vert \rightarrow +\infty . \end{aligned}$$

For this purpose, we divide our proof into two steps. In the first step, we prove inequality for the \(\liminf \). At the second step, we show inequality for the \(\limsup \).

Step 1. the lower estimate We show that

$$\begin{aligned} \liminf _{\vert t-s\vert \rightarrow +\infty }\frac{d(\varGamma _{t},\varGamma _{s})}{\vert t-s\vert }\ge c_{f}. \end{aligned}$$
(29)

We assume that (29) does not hold, then one has

$$\begin{aligned} \liminf _{\vert t-s\vert \rightarrow +\infty }\frac{d(\varGamma _{t},\varGamma _{s})}{\vert t-s\vert }< c_{f}-2\varepsilon \end{aligned}$$
(30)

for some \(\varepsilon >0\) small enough. Thus, there exist two sequences \((t_{k})_{k\in \mathbb {N}}\) and \((s_{k})_{k\in \mathbb {N}}\) such that \(\vert t_{k}-s_{k}\vert \rightarrow +\infty \) as \(k\rightarrow +\infty \) and

$$\begin{aligned} d(\varGamma _{t_{k}},\varGamma _{s_{k}})<(c_{f}-2\varepsilon )\vert t_{k}-s_{k}\vert \quad \mathrm{for}~k~\mathrm{large~enough}. \end{aligned}$$

Without loss of generality, we assume that \(t_{k}<s_{k}\) for all \(k\in \mathbb {N}\). The definition of the distance \(d(\varGamma _{t_{k}},\varGamma _{s_{k}})\) implies that there exist two sequences \((x_{k})_{k\in \mathbb {N}}\) and \((z_{k})_{k\in \mathbb {N}}\) in \(\mathbb {R}^{N}\) such that

$$\begin{aligned} x_{k}\in \varGamma _{t_{k}},~z_{k}\in \varGamma _{s_{k}}~\mathrm{and}~\vert x_{k}-z_{k}\vert <(c_{f}-2\varepsilon )(s_{k}-t_{k})\quad \mathrm{for}~k~\mathrm{large~enough}. \end{aligned}$$

First of all, by (28), there exists a sequence \((y_{k}^{+})_{k\in \mathbb {N}}\) of points in \(\mathbb {R}^{N}\) such that

$$\begin{aligned} y_{k}^{+}\in \varOmega _{t_{k}}^{+},~\vert x_{k}-y_{k}^{+}\vert \le r~~\mathrm{and}~~d(y_{k}^{+},\varGamma _{t_{k}})\ge 2R~~\mathrm{for~all}~k\in \mathbb {N}. \end{aligned}$$

Thus, for every \(k\in \mathbb {N}\) and \(y\in B(y_{k}^{+},R)\), one has \(y\in \varOmega _{t_{k}}^{+}\) and \(d(y,\varGamma _{t_{k}})\ge R\ge M\), whence \(u(t_{k},y)\ge \beta \) from (27). By (11), one has

$$\begin{aligned} u(t_{k},x)\ge v_{R}^{f}(0,x-y_{k}^{+}),\quad x\in \mathbb {R}^{N}. \end{aligned}$$

Thus the maximum principle yields

$$\begin{aligned} u(t,x)\ge v_{R}^{f}(t-t_{k},x-y_{k}^{+})\quad \mathrm{for~all}~t>t_{k}~\mathrm{and}~x\in \mathbb {R}^{N}. \end{aligned}$$

Let \(T_{\varepsilon }\) be defined as in Lemma 1, thus Lemma 1 yields that for every \(k\in \mathbb {N}\),

$$\begin{aligned} u(t,x)\ge \beta \quad \mathrm{for~all}~t\ge t_{k}+T_{\varepsilon }~\mathrm{and}~\vert x-y_{k}^{+}\vert \le (c_{f}-\varepsilon )(t-t_{k}). \end{aligned}$$
(31)

Next, it follows from (28) that there exists a sequence \((y_{k}^{-})_{k\in \mathbb {N}}\) of points in \(\mathbb {R}^{N}\) such that

$$\begin{aligned} y_{k}^{-}\in \varOmega _{s_{k}}^{-},~\vert z_{k}-y_{k}^{-}\vert \le r~~\mathrm{and}~~d(y_{k}^{-},\varGamma _{s_{k}})\ge 2R\ge M~~\mathrm{for~all}~k\in \mathbb {N}. \end{aligned}$$

Property (27) implies that

$$\begin{aligned} u(s_{k},y_{k}^{-})\le \alpha _{\varepsilon }\quad \mathrm{for~all}~k\in \mathbb {N}. \end{aligned}$$
(32)

Finally, notice that for all \(k\in \mathbb {N}\),

$$\begin{aligned} \vert y_{k}^{-}-y_{k}^{+}\vert \le \vert y_{k}^{-}-z_{k}\vert +\vert z_{k}-x_{k}\vert +\vert x_{k}-y_{k}^{+}\vert \le r+(c_{f}-2\varepsilon )(s_{k}-t_{k})+r \end{aligned}$$

Thus, it follows from \(s_{k}-t_{k}\rightarrow +\infty \) as \(k\rightarrow +\infty \) that \(s_{k}\ge t_{k}+T_{\varepsilon }\) for k large enough and

$$\begin{aligned} \vert y_{k}^{-}-y_{k}^{+}\vert \le (c_{f}-\varepsilon )(s_{k}-t_{k})\quad \mathrm{for}~k~\mathrm{large~enough}. \end{aligned}$$

Choose \(t=s_{k}\) and \(x=y_{k}^{-}\) in (31) for k large enough. Thus,

$$\begin{aligned} u(s_{k},y_{k}^{-})\ge \beta \quad \mathrm{for}~k~\mathrm{large~enough}. \end{aligned}$$

But \(\alpha _{\varepsilon }<\beta \) contradicting (32). Therefore, the assumption (30) cannot hold. That is,

$$\begin{aligned} \liminf _{\vert t-s\vert \rightarrow +\infty }\frac{d(\varGamma _{t},\varGamma _{s})}{\vert t-s\vert }\ge c_{f}. \end{aligned}$$

Step 2: the upper estimate We show that

$$\begin{aligned} \limsup _{\vert t-s\vert \rightarrow +\infty }\frac{d(\varGamma _{t},\varGamma _{s})}{\vert t-s\vert }\le c_{f}. \end{aligned}$$
(33)

Let us assume by contradiction that

$$\begin{aligned} \limsup _{\vert t-s\vert \rightarrow +\infty }\frac{d(\varGamma _{t},\varGamma _{s})}{\vert t-s\vert }> c_{f}+3\varepsilon \end{aligned}$$
(34)

for some \(\varepsilon >0\) small enough. Then there exist two sequences \((t_{k})_{k\in \mathbb {N}}\) and \((s_{k})_{k\in \mathbb {N}}\) of real numbers such that \(\vert t_{k}-s_{k}\vert \rightarrow +\infty \) as \(k\rightarrow +\infty \) and

$$\begin{aligned} d(\varGamma _{t_{k}},\varGamma _{s_{k}})>(c_{f}+3\varepsilon )\vert t_{k}-s_{k}\vert \quad \mathrm{for}~k~\mathrm{large~enough}. \end{aligned}$$

Without loss of generality, one can assume that \(t_{k}<s_{k}\) for all \(k\in \mathbb {N}\). For each \(k\in \mathbb {N}\), pick a point \(z_{k}\) on \(\varGamma _{s_{k}}\). It follows from (28) that there are two sequences \((y_{k}^{\pm })_{k\in \mathbb {N}}\) of points in \(\mathbb {R}^{N}\) such that

$$\begin{aligned} y_{k}^{\pm }\in \varOmega _{s_{k}}^{\pm },~\vert z_{k}-y_{k}^{\pm }\vert \le r~\mathrm{and}~d(y_{k}^{\pm },\varGamma _{s_{k}})\ge M\quad \mathrm{for~all}~k\in \mathbb {N}. \end{aligned}$$

Thus, by (27), one has

$$\begin{aligned} 0<u(s_{k},y_{k}^{-})\le \alpha _{\varepsilon }<3\alpha _{\varepsilon }<\theta<\beta \le u(s_{k},y_{k}^{+})<1\quad \mathrm{for~all}~k\in \mathbb {N}. \end{aligned}$$
(35)

It follows from \(d(z_{k},\varGamma _{t_{k}})>(c_{f}+3\varepsilon )( s_{k}-t_{k})>0\) for k large enough that there holds

$$\begin{aligned} \mathrm{either}~B(z_{k},(c_{f}+3\varepsilon )( s_{k}-t_{k}))\subset \varOmega _{t_{k}}^{+} ~\mathrm{or}~B(z_{k},(c_{f}+3\varepsilon )( s_{k}-t_{k}))\subset \varOmega _{t_{k}}^{-}. \end{aligned}$$

We claim that \(B(z_{k},(c_{f}+3\varepsilon )( s_{k}-t_{k}))\subset \varOmega _{t_{k}}^{-}\) for k large enough. If not, up to extraction of a subsequence,

$$\begin{aligned} B(z_{k},(c_{f}+3\varepsilon )( s_{k}-t_{k}))\subset \varOmega _{t_{k}}^{+}\quad \mathrm{for~all}~k~\mathrm{large~enough}. \end{aligned}$$
(36)

Since \(s_{k}-t_{k}\rightarrow +\infty \) as \(k\rightarrow +\infty \), then for k large enough,

$$\begin{aligned} B(z_{k},R)\subset \varOmega _{t_{k}}^{+}~\mathrm{and}~d(y,\varGamma _{t_{k}})\ge M\quad \mathrm{for~all~} y\in B(z_{k},R), \end{aligned}$$

recall that \(R>0\) is defined as in Lemma 1. Thus for k large enough,

$$\begin{aligned} u(t_{k},y)\ge \beta \quad \mathrm{for~all}~y\in B(z_{k},R), \end{aligned}$$

from (27). It follows from (11) that one has

$$\begin{aligned} u(t_{k},x)\ge v_{R}^{f}(0,x-z_{k}),\quad \mathrm{for~all~}x\in \mathbb {R}^{N}. \end{aligned}$$

By the maximum principle, one gets

$$\begin{aligned} u(t,x)\ge v_{R}^{f}(t-t_{k},x-z_{k})\quad \mathrm{for~all}~t>t_{k}~\mathrm{and}~x\in \mathbb {R}^{N}. \end{aligned}$$
(37)

Let \(T_{\varepsilon '}>0\) be defined as in Lemma 1 with \(\varepsilon '=\frac{c_{f}}{2}\). The inequality (37) and Lemma 1 yield that, for k large enough,

$$\begin{aligned} u(t,x)\ge \beta ~~\mathrm{for~all}~t\ge t_{k}+T_{\varepsilon '}~\mathrm{and}~\vert x-z_{k}\vert \le (c_{f}-\varepsilon ')(t-t_{k})=\frac{c_{f}}{2}(t-t_{k}). \end{aligned}$$

Since \(c_{f}>0\) and \(s_{k}-t_{k}\rightarrow +\infty \) as \(k\rightarrow +\infty \), then for k large enough,

$$\begin{aligned} s_{k}\ge t_{k}+T_{\varepsilon '}~\mathrm{and}~ \vert y_{k}^{-}-z_{k}\vert \le r\le \frac{c_{f}}{2}(s_{k}-t_{k}). \end{aligned}$$

Thus, the previous inequality imples \(u(s_{k},y_{k}^{-})\ge \beta \) for k large enough. This is in contradiction with (35). Whence for k large enough,

$$\begin{aligned} B(z_{k},(c_{f}+3\varepsilon )( s_{k}-t_{k}))\subset \varOmega _{t_{k}}^{-}. \end{aligned}$$

Since \(s_{k}-t_{k}\rightarrow +\infty \) as \(k\rightarrow +\infty \), then for k large enough,

$$\begin{aligned} B(z_{k},(c_{f}+2\varepsilon )( s_{k}-t_{k}))\subset \varOmega _{t_{k}}^{-}~\mathrm{and}~ d(y,\varGamma _{t_{k}})\ge M~\mathrm{for~all}~y\in B(z_{k},(c_{f}+2\varepsilon )( s_{k}-t_{k})). \end{aligned}$$

Thus by (27), one has

$$\begin{aligned} u(t_{k},y)\le \alpha _{\varepsilon }~~\mathrm{for~all}~y\in B(z_{k},(c_{f}+2\varepsilon )( s_{k}-t_{k})). \end{aligned}$$

It follows from the definition of \(w_{(c_{f}+2\varepsilon )( s_{k}-t_{k})}\) (as in Lemma 1) that for k large enough,

$$\begin{aligned} u(t_{k},x)\le w_{(c_{f}+2\varepsilon )( s_{k}-t_{k})}(0,x-z_{k})~~\mathrm{for~all}~ x\in \mathbb {R}^{N}. \end{aligned}$$

Thus the maximum principle implies

$$\begin{aligned} u(t,x)\le w_{(c_{f}+2\varepsilon )( s_{k}-t_{k})}(t-t_{k},x-z_{k})\quad \mathrm{for~all}~t>t_{k}~\mathrm{and}~x\in \mathbb {R}^{N}. \end{aligned}$$

Since \(s_{k}-t_{k}\rightarrow +\infty \) as \(k\rightarrow +\infty \), then for k large enough, one has \((c_{f}+2\varepsilon )(s_{k}-t_{k})\ge R_{\varepsilon }\) and

$$\begin{aligned} \begin{aligned}&T_{\varepsilon }\le s_{k}-t_{k}\le \frac{(c_{f}+2\varepsilon )( s_{k}-t_{k})}{c_{f}+\varepsilon },\\&\vert y_{k}^{+}-z_{k}\vert \le r\le \varepsilon (s_{k}-t_{k})=(c_{f}+2\varepsilon )( s_{k}-t_{k})-(c_{f}+\varepsilon )( s_{k}-t_{k}), \end{aligned} \end{aligned}$$

where \(T_{\varepsilon }>0\) and \(R_{\varepsilon }\ge (c_{f}+\varepsilon )T_{\varepsilon }>0\) are given by Lemma 2 so that (16) is valid for all \(R\ge R_{\varepsilon }\). Choose \(R=(c_{f}+2\varepsilon )( s_{k}-t_{k})\), \(t=s_{k}-t_{k}\) and \(x=y_{k}^{+}-z_{k}\) in (16) for k large enough, one can obtain

$$\begin{aligned} u(s_{k},y_{k}^{+})\le w_{(c_{f}+2\varepsilon )( s_{k}-t_{k})}(s_{k}-t_{k},y_{k}^{+}-z_{k})\le 3\alpha _{\varepsilon }. \end{aligned}$$

This is in contradiction with (35). Whence the conclusion (33) follows.

Combining with the Step 1 and Step 2, the proof of Theorem 1 is thereby complete.

3 Almost-Planar Fronts

In this section, we characterize the planar fronts \(\phi _{f}(x\cdot e-c_{f}t)\) for Eq. (1) among the more general class of almost-planar fronts. The proof of Theorem 2 mainly uses the one-dimensional stability of the planar front \(\phi _{f}\) [18] and the parabolic Liouville type result of Berestycki and Hamel [3, Theorem 3.1]. Before the proof, we first give some auxiliary lemmas.

Lemma 3

Let \(u:\mathbb {R}\times \mathbb {R}^{N}\rightarrow [0,1]\) be a solution of (1) for which there are a real number \(t_{0}\in \mathbb {R}\) and a unit vector \(e\in \mathbb {S}^{N-1}\) such that

$$\begin{aligned} \sup _{x\in \mathbb {R}^{N},x\cdot e\ge A}u(t_{0},x)\rightarrow 0\quad \left( resp.\inf _{x\in \mathbb {R}^{N},x\cdot e\le -A}u(t_{0},x)\rightarrow 1 \right) \quad as~A\rightarrow +\infty . \end{aligned}$$
(38)

Then property (38) holds at every time \(t_{1}>t_{0}\) with the same vector e.

Proof

Since for the case \(\inf \limits _{x\in \mathbb {R}^{N},x\cdot e\le -A}u(t_{0},x)\rightarrow 1\) as \(A\rightarrow +\infty \), the proof of Lemma 3 is similar to [11, Lemma 3.1], we only give the proof for the case \(\sup \limits _{x\in \mathbb {R}^{N},x\cdot e\ge A}u(t_{0},x)\rightarrow 0\) as \(A\rightarrow +\infty \).

For any \(\delta \in (0,1)\), let \(v^{\delta }\) be the solution of the following one-dimensional Cauchy problem

$$\begin{aligned} \begin{aligned}&v^{\delta }_{t}=v^{\delta }_{yy}+f(v^{\delta }),\quad t>0,~y\in \mathbb {R},\\&v^{\delta }(0,y)= {\left\{ \begin{array}{ll} 1,\quad y\le 0,\\ \delta ,\quad y>0. \end{array}\right. } \end{aligned} \end{aligned}$$

Let \(\rho ^{\delta }:\mathbb {R}\rightarrow (0,1)\) be the solution of the following ordinary differential equation

$$\begin{aligned} \begin{aligned}&\left( \rho ^{\delta }\right) ^{\prime }(t)=f(\rho ^{\delta }(t)),\quad t>0,\\&\rho ^{\delta }(0)=\delta .\quad \quad \quad \quad \quad \quad \quad \end{aligned} \end{aligned}$$

Then by the maximum principle, one has

$$\begin{aligned} 0\le \rho ^{\delta }(t)\le v^{\delta }(t,y)\le 1,\quad t\ge 0,\quad y\in \mathbb {R}. \end{aligned}$$

Thus

$$\begin{aligned} 0\le v^{\delta }(t,x)-\rho ^{\delta }(t)\le \frac{e^{Lt}}{(4\pi t)^{\frac{N}{2}}}\int _{y\le 0}e^{-\frac{\vert x-y\vert ^{2}}{4t}}dy, \end{aligned}$$

where \(L=\max \limits _{u\in [0,1]}\vert f'(u)\vert \). Then the maximum principle and standard parabolic estimates imply that for each \(t>0\), \(v^{\delta }(t,\cdot )\) is decreasing in \(\mathbb {R}\), \(v^{\delta }(t,-\infty )=1\) and \(v^{\delta }(t,+\infty )=\rho ^{\delta }(t)\).

Assume that \(\sup \limits _{x\in \mathbb {R}^{N},x\cdot e\ge A}u(t_{0},x)\rightarrow 0\) as \(A\rightarrow +\infty \). Let \(\varepsilon \in (0,\theta )\) be arbitrary. Then there exists a constant M such that

$$\begin{aligned} u(t_{0},x)\le v^{\varepsilon }(0,x\cdot e-M),\quad x\in \mathbb {R}^{N}. \end{aligned}$$

Thus it follows from the maximum principle that

$$\begin{aligned} u(t_{1},x)\le v^{\varepsilon }(t_{1}-t_{0},x\cdot e-M),\quad t_{1}>t_{0},~ x\in \mathbb {R}^{N}, \end{aligned}$$

and whence

$$\begin{aligned} \limsup _{A\rightarrow +\infty }\left( \sup \limits _{x\in \mathbb {R}^{N},x\cdot e\ge A}u(t_{1},x)\right) \le v^{\varepsilon }(t_{1}-t_{0},+\infty )=\rho ^{\varepsilon }(t_{1}-t_{0}),\quad t_{1}>t_{0}. \end{aligned}$$

Since \(\varepsilon \in (0,\theta )\), f is \(C^{1}\) on [0, 1] and \(f=0\) on \([0,\theta ]\), then the existence and uniqueness of solution of the ordinary differential equation yield that \(\rho ^{\varepsilon }(t)\equiv \varepsilon \) for all \(t\ge 0\). Therefore, one has

$$\begin{aligned} 0\le \limsup _{A\rightarrow +\infty }\left( \sup \limits _{x\in \mathbb {R}^{N},x\cdot e\ge A}u(t_{1},x)\right) \le \rho ^{\varepsilon }(t_{1}-t_{0})=\varepsilon \rightarrow 0\quad \mathrm{as}~\varepsilon \rightarrow 0. \end{aligned}$$

This completes the proof. \(\square \)

The following corollary can be obtained immediately from Lemma 3.

Corollary 1

Let \(u:\mathbb {R}\times \mathbb {R}^{N}\rightarrow [0,1]\) be a solution of (1) such that, for every time \(t\in \mathbb {R}\), there is a unit vector \(e_{t}\in \mathbb {S}^{N-1}\) such that

$$\begin{aligned} \inf _{x\in \mathbb {R}^{N},x\cdot e_{t}\le -A}u(t,x)\rightarrow 1\quad and\quad \sup _{x\in \mathbb {R}^{N},x\cdot e_{t}\ge A}u(t,x)\rightarrow 0\quad as~A\rightarrow +\infty . \end{aligned}$$
(39)

Then \(e_{t}=e\) is independent of time t.

Let u be an almost-planar transition front connecting 0 and 1, in the sense of Definition 2, for problem (1). That is, there exist some families \((e_{t})_{t\in \mathbb {R}}\) in \(\mathbb {S}^{N-1}\) and \((\xi _{t})_{t\in \mathbb {R}}\) in \(\mathbb {R}\) such that

$$\begin{aligned} \varGamma _{t}=\left\{ x\in \mathbb {R}^{N}\big \vert x\cdot e_{t}=\xi _{t}\right\} \end{aligned}$$

for every \(t\in \mathbb {R}\). Up to changing \(e_{t}\) into \(-e_{t}\), (5) and Definition 1 yields that (39) holds for every \(t\in \mathbb {R}\). It follows from Corollary 1 that \(e_{t}=e\) is a constant vector, whence

$$\begin{aligned} \varOmega _{t}^{+}=\left\{ x\in \mathbb {R}^{N}\big \vert x\cdot e<\xi _{t}\right\} \quad \mathrm{and}\quad \varOmega _{t}^{-}=\left\{ x\in \mathbb {R}^{N}\big \vert x\cdot e>\xi _{t}\right\} \end{aligned}$$
(40)

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

In Sect. 2, we have already proved that any transition front connecting equilibrium points 0 and 1 has a global mean speed \(c_{f}\). Here, for almost planar fronts, one has that

$$\begin{aligned} \frac{\vert \xi _{t}-\xi _{s}\vert }{\vert t-s\vert }\rightarrow c_{f}\quad \mathrm{as}~\vert t-s\vert \rightarrow +\infty . \end{aligned}$$

Then for any \(\gamma \in (0,1)\), there exists a constant \(K>0\) large enough such that

$$\begin{aligned} \gamma c_{f}\vert t-s\vert \le \vert \xi _{t}-\xi _{s}\vert \le 2c_{f}\vert t-s\vert \quad \mathrm{for}~\vert t-s\vert \ge K. \end{aligned}$$
(41)

For \(n\in \mathbb {Z}\), we define \(\widetilde{\xi }_{t}\) such that

$$\begin{aligned} \widetilde{\xi }_{t}= {\left\{ \begin{array}{ll} \xi _{t},\quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \quad \, t=nK,\\ \xi _{nK}+\frac{\xi _{(n+1)K}-\xi _{nK}}{K}(t-nK),\quad nK\le t\le (n+1)K. \end{array}\right. } \end{aligned}$$

It follows from (9) that one has

$$\begin{aligned} \forall ~\sigma >0,~~\sup _{(t,s)\in \mathbb {R}^{2},\vert t-s\vert \le \sigma }\vert \xi _{t}-\xi _{s}\vert <+\infty . \end{aligned}$$

Thus, one gets

$$\begin{aligned} \sup _{t\in \mathbb {R}}\vert \widetilde{\xi }_{t}-\xi _{t}\vert <+\infty . \end{aligned}$$
(42)

Moreover, one has

$$\begin{aligned} \gamma c_{f}\le \frac{\xi _{(n+1)K}-\xi _{nK}}{K}\le 2c_{f}. \end{aligned}$$

Now we mollify the function \(\widetilde{\xi }_{t}\) to make it smooth. Define \(\eta \in C^{\infty }(\mathbb {R})\) by

$$\begin{aligned} \eta (z)= {\left\{ \begin{array}{ll} C\exp \left( \frac{1}{\vert z\vert ^{2}-1}\right) \quad &{} \mathrm{if}~~\vert z\vert <1,\\ 0\quad &{} \mathrm{if}~~\vert z\vert \ge 1,\\ \end{array}\right. } \end{aligned}$$

where the constant \(C>0\) is selected so that \(\int _{\mathbb {R}}\eta dz=1\). For each \(\epsilon >0\), set \(\eta _{\epsilon }(z)=\frac{1}{^{\epsilon }}\eta \left( \frac{z}{\epsilon }\right) \). Let

$$\begin{aligned} \xi _{t}^{\epsilon }=\widetilde{\xi }_{t}*\eta _{\epsilon }=\int _{-\epsilon }^{\epsilon } \eta _{\epsilon }(z)\widetilde{\xi }_{t-z}dz,\quad t\in \mathbb {R} \end{aligned}$$

such that

$$\begin{aligned} \sup _{t\in \mathbb {R}}\vert \xi _{t}^{\epsilon }-\widetilde{\xi }_{t}\vert \le 1\quad \mathrm{and}\quad \gamma c_{f}\le \frac{d\xi _{t}^{\epsilon }}{dt}\le 2c_{f}\quad \mathrm{for}~ t\in \mathbb {R}. \end{aligned}$$
(43)

Whence, (42) and (43) yield

$$\begin{aligned} \sup _{t\in \mathbb {R}}\vert \xi _{t}^{\epsilon }-\xi _{t}\vert \le +\infty , \end{aligned}$$

and hence, u(tx) is still an almost-planar front with sets

$$\begin{aligned} \widetilde{\varGamma }_{t}=\left\{ x\in \mathbb {R}^{N}\big \vert x\cdot e=\xi _{t}^{\epsilon }\right\} \end{aligned}$$

and

$$\begin{aligned} \widetilde{\varOmega }_{t}^{+}=\left\{ x\in \mathbb {R}^{N}\big \vert x\cdot e<\xi _{t}^{\epsilon }\right\} ,\quad \widetilde{\varOmega }_{t}^{-}=\left\{ x\in \mathbb {R}^{N}\big \vert x\cdot e>\xi _{t}^{\epsilon }\right\} . \end{aligned}$$

from Remark 1. Let \(\alpha \) and \(\beta \) be two given real numbers such that

$$\begin{aligned} 0<\alpha<\theta<\beta <1, \end{aligned}$$
(44)

where we recall that \(\theta \) is defined in (2). By the Definition 1, there is \(M>0\) such that

$$\begin{aligned} \forall ~(t,x)\in \mathbb {R}\times \mathbb {R}^{N},\quad {\left\{ \begin{array}{ll} x\cdot e-\xi _{t}^{\epsilon }\le -M\Rightarrow \beta \le u(t,x)<1,\\ x\cdot e-\xi _{t}^{\epsilon }\ge M\Rightarrow 0<u(t,x)\le \alpha . \end{array}\right. } \end{aligned}$$
(45)

Lemma 4

For any \(\gamma \in (0,1)\),

$$\begin{aligned} u(t,x)\le \theta e^{-\gamma c_{f}(x\cdot e-\xi _{t}^{\epsilon }-M)} \end{aligned}$$
(46)

in \(\varSigma =\left\{ (t,x)\in \mathbb {R}\times \mathbb {R}^{N}\big \vert x\cdot e-\xi _{t}^{\epsilon }\ge M\right\} \).

Proof

Let \(\bar{u}(t,x)=\theta e^{-\gamma c_{f}(x\cdot e-\xi _{t}^{\epsilon }-M)}\). On \(\partial \varSigma =\left\{ (t,x)\in \mathbb {R}\times \mathbb {R}^{N}\big \vert x\cdot e-\xi _{t}^{\epsilon }= M\right\} \), it follows from (44) and (45) that

$$\begin{aligned} \bar{u}(t,x)=\theta >\alpha \ge u(t,x). \end{aligned}$$

Define

$$\begin{aligned} \varepsilon _{*}=\inf \left\{ \varepsilon >0\big \vert u-\varepsilon \le \bar{u}~\mathrm{in}~\varSigma \right\} . \end{aligned}$$

Since u is bounded, \(\varepsilon _{*}\) is a well-defined real number and \(\varepsilon _{*}\ge 0\). Furthermore, one has

$$\begin{aligned} w:=\bar{u}-(u-\varepsilon _{*})\ge 0\quad \mathrm{in}~\varSigma . \end{aligned}$$

In particular,

$$\begin{aligned} w>\varepsilon _{*}\quad \mathrm{on}~\partial \varSigma . \end{aligned}$$
(47)

One only has to prove that \(\varepsilon _{*}=0\).

Assume by contradiction that \(\varepsilon _{*}>0\). Then there exists a sequence \((\varepsilon _{n})_{n\in \mathbb {N}}\) of positive real numbers and a sequence of points \((t_{n},x_{n})_{n\in \mathbb {N}}\) in \(\varSigma \) such that

$$\begin{aligned} \varepsilon _{n}\rightarrow \varepsilon _{*}~\mathrm{as}~n\rightarrow +\infty ~~\mathrm{and}~\bar{u}(t_{n},x_{n})< u(t_{n},x_{n})-\varepsilon _{n}~\mathrm{for~all}~n\in \mathbb {N}. \end{aligned}$$

We claim that the sequence \((x_{n}\cdot e-\xi _{t_{n}}^{\epsilon })_{n\in \mathbb {N}}\) is bounded. Assume not, up to extraction of some sequence, one has

$$\begin{aligned} x_{n}\cdot e-\xi _{t_{n}}^{\epsilon }\rightarrow +\infty ,~\mathrm{and~then}~u(t_{n},x_{n})\rightarrow 0\quad \mathrm{as}~n\rightarrow +\infty . \end{aligned}$$

But

$$\begin{aligned} u(t_{n},x_{n})>\bar{u}(t_{n},x_{n})+\varepsilon _{n} \ge \varepsilon _{n}\rightarrow \varepsilon _{*}>0\quad \mathrm{as}~n\rightarrow +\infty . \end{aligned}$$

This gives a contradiction. Thus, the sequence \((x_{n}\cdot e-\xi _{t_{n}}^{\epsilon })_{n\in \mathbb {N}}\) is bounded.

It follows from (9) that for any \(\sigma >0\), there holds

$$\begin{aligned} \sup \{d(\widetilde{\varGamma }_{t},\widetilde{\varGamma }_{s}),t,s\in \mathbb {R},\vert t-s\vert \le \sigma \}<+\infty . \end{aligned}$$

Since \((\widetilde{\varGamma }_t)_{t\in \mathbb {R}}\) are all parallel hyperplanes, it then follows that for any fix \(\tau >0\), there exists a sequence \((\tilde{x}_{n})_{n\in \mathbb {N}}\) such that

$$\begin{aligned} \tilde{x}_{n}\in \widetilde{\varGamma }_{t_{n}-\tau }~\mathrm{for~all}~n\in \mathbb {N}~\mathrm{and}~\sup \{d(x_{n},\tilde{x}_{n})\}<+\infty . \end{aligned}$$

By (6), there exist \(r>0\) and a sequence \((y_{n})_{n\in \mathbb {N}}\) such that

$$\begin{aligned} d(\tilde{x}_{n},y_{n})\le r\quad \mathrm{and}\quad y_{n}\cdot e-\xi _{t_{n}-\tau }^{\epsilon }\ge M~\mathrm{for~all}~n\in \mathbb {N}. \end{aligned}$$

Then there exists a sequence \((z_{n})_{n\in \mathbb {N}}\) such that

$$\begin{aligned} z_{n}\in \overline{\widetilde{\varOmega }_{t_{n}-\tau }^{-}}\quad \mathrm{and}\quad M=z_{n}\cdot e-\xi _{t_{n}-\tau }^{\epsilon }=y_{n}\cdot e-\xi _{t_{n}-\tau }^{\epsilon }-d(y_{n},z_{n})~\mathrm{for~all}~n\in \mathbb {N}. \end{aligned}$$
(48)

Since \(d(y_{n},z_{n})\le y_{n}\cdot e-\xi _{t_{n}-\tau }^{\epsilon }\le d(\tilde{x}_{n},y_{n})\le r\) and since the sequence \((d(\tilde{x}_{n},x_{n}))_{n\in \mathbb {N}}\) is bounded, then the sequence \((d(x_{n},z_{n}))_{n\in \mathbb {N}}\) is bounded.

Choose \(\rho >0\) so that

$$\begin{aligned} \rho \Vert (\bar{u}-u)_{t}\Vert _{L^{\infty }(\mathbb {R}\times \varSigma )} +2\rho \Vert \nabla _{x}(\bar{u}-u)\Vert _{L^{\infty }(\mathbb {R}\times \varSigma )}<\varepsilon _{*}, \end{aligned}$$
(49)

which is possible since \(\bar{u}\) and u have bounded derivatives. Choose \(K\in \mathbb {N}\setminus \{0\}\) so that

$$\begin{aligned} K\rho \ge \max \left( \tau ,\sup \{d(x_{n},z_{n})\big \vert n\in \mathbb {N}\}\right) . \end{aligned}$$
(50)

For each \(n\in \mathbb {N}\), then there exists a sequence of points \((X_{n,0},X_{n,1},\ldots ,X_{n,K})\) in \(\varSigma \) such that

$$\begin{aligned} X_{n,0}=x_{n},~X_{n,K}=z_{n}~\mathrm{and}~d(X_{n,i},X_{n,i+1})\le \rho ~\mathrm{for~each}~0\le i\le K-1. \end{aligned}$$

For each \(n\in \mathbb {N}\) and \(0\le i\le K-1\), set

$$\begin{aligned} E_{n,i}=\left[ t_{n}-\frac{i+1}{K}\tau ,t_{n} -\frac{i}{K}\tau \right] \times \overline{B(X_{n,i},2\rho )}. \end{aligned}$$

Since \(w(t_{n},x_{n})\rightarrow 0\) as \(n\rightarrow +\infty \), (49) and (50) yield that \(w<\varepsilon _{*}\) in \(E_{n,0}\) for large n. It follows from (47) and the connectivity of \(E_{n,0}\) that \(E_{n,0}\subset \varSigma \) for large n.

By the definition of \(\varSigma \) and \(\bar{u}\), one has \(0\le \bar{u}\le \theta \) in \(\varSigma \). Then from (2) and (43), one has

$$\begin{aligned} \bar{u}_{t}-\varDelta \bar{u}-f(\bar{u})=\,&\gamma c_{f}\bar{u}\frac{d\xi _{t}^{\epsilon }}{dt}-(\gamma c_{f})^{2}\bar{u}\\ =\,&\bar{u}\gamma c_{f}\left[ \frac{d\xi _{t}^{\epsilon }}{dt}-\gamma c_{f}\right] \\ \ge \,&0 \end{aligned}$$

in \(\varSigma \). On the other hand, \(u-\varepsilon _{*}<u\le \alpha <\theta \) in \(\varSigma \). Assumption (2) implies that \(u-\varepsilon _{*}\) is a subsolution of (1) in \(\varSigma \). Since f is of class \(C^{1}\), the function w satisfies inequations of the type

$$\begin{aligned} w_{t}\ge \varDelta w+b(t,x)w\quad \mathrm{in}~E_{n,0} \end{aligned}$$

for n large enough, where the sequence \((\Vert b\Vert _{L^{\infty }(E_{n,0})})_{n\in \mathbb {N}}\) is bounded. Since \(w(t_{n},X_{n,0})=w(t_{n},x_{n})\rightarrow 0\) as \(n\rightarrow +\infty \), it follows from the linear parabolic estimates that

$$\begin{aligned} w\left( t_{n}-\frac{\tau }{K},X_{n,1}\right) \rightarrow 0\quad \mathrm{as}~n\rightarrow +\infty . \end{aligned}$$

An immediate induction yields \(w\left( t_{n}-\frac{i\tau }{K},X_{n,i}\right) \rightarrow 0\) as \(n\rightarrow +\infty \) for each \(i=1,\ldots ,K\). In particular, for \(i=K\),

$$\begin{aligned} w\left( t_{n}-\tau ,z_{n}\right) \rightarrow 0\quad \mathrm{as}~n\rightarrow +\infty . \end{aligned}$$

But \(z_{n}\in \overline{\widetilde{\varOmega }_{t_{n}-\tau }^{-}}\) and \(z_{n}\cdot e-\xi _{t_{n}-\tau }^{\epsilon }=M\) for all \(n\in \mathbb {N}\). As a consequence, for all \(n\in \mathbb {N}\), \(w\left( t_{n}-\tau ,z_{n}\right) >\varepsilon _{*}\) from (47).

One has reached a contradiction, which means that \(\varepsilon _{*}=0\). Thus,

$$\begin{aligned} u(t,x)\le \theta e^{-\gamma c_{f}(x\cdot e-\xi _{t}^{\epsilon }-M)} \end{aligned}$$

for all \((t,x)\in \varSigma \). This completes the proof. \(\square \)

Proof of Theorem 2

For any fixed \(\gamma \in (0,1)\), let \(\underline{v}_{\beta }\) and \(\bar{v}_{\alpha }\) be the solution of the one-dimensional Cauchy problem

$$\begin{aligned} v_{t}=v_{yy}+f(v),\quad t>0,~y\in \mathbb {R} \end{aligned}$$
(51)

with initial condition

$$\begin{aligned} \underline{v}_{\beta }(0,y)\in C(\mathbb {R},[0,\beta ])\quad \mathrm{and}\quad \underline{v}_{\beta }(0,y)= {\left\{ \begin{array}{ll} \beta \quad \mathrm{if}~y\le -1,\\ 0\quad \mathrm{if}~y\ge 0,\quad \end{array}\right. } \end{aligned}$$
(52)

and \(\bar{v}_{\alpha }(0,y)\in C(\mathbb {R},[0,1])\),

$$\begin{aligned} \bar{v}_{\alpha }(0,y)\ge \alpha ~\mathrm{on}~[-1,1]\quad \mathrm{and}\quad \bar{v}_{\alpha }(0,y)= {\left\{ \begin{array}{ll} 1\quad \mathrm{if}~y\le 0,\\ \theta e^{-\gamma c_{f}y}\quad \mathrm{if}~y\ge 1, \end{array}\right. } \end{aligned}$$
(53)

respectively. It follows from (45) and Lemma 4 that for every \(t_{0}\in \mathbb {R}\) and \(x\in \mathbb {R}^{N}\),

$$\begin{aligned} \underline{v}_{\beta }(0,x\cdot e-\xi _{t_{0}}^{\epsilon }+M)\le u(t_{0},x)\le \bar{v}_{\alpha }(0,x\cdot e-\xi _{t_{0}}^{\epsilon }-M). \end{aligned}$$

Thus,

$$\begin{aligned} \underline{v}_{\beta }(t-t_{0},x\cdot e-\xi _{t_{0}}^{\epsilon }+M)\le u(t,x)\le \bar{v}_{\alpha }(t-t_{0},x\cdot e-\xi _{t_{0}}^{\epsilon }-M). \end{aligned}$$
(54)

for all \(t>t_{0}\) and \(x\in \mathbb {R}^{N}\), from the maximum principle. By [18], there exist two constants \(\underline{\omega }>0\) and \(\bar{\omega }>0\) such that

$$\begin{aligned} \left| \frac{\underline{v}_{\beta }(s,y)-\phi _{f}(y-c_{f}s+\underline{\xi })}{\phi _{f}^{\gamma }(y-c_{f}s+\underline{\xi })}\right| \le \underline{A}e^{-\underline{\omega } t},\quad s\ge 0,\quad y\in \mathbb {R}, \end{aligned}$$

and

$$\begin{aligned} \left| \frac{\bar{v}_{\alpha }(s,y)-\phi _{f}(y-c_{f}s+\bar{\xi })}{\phi _{f}^{\gamma }(y-c_{f}s+\bar{\xi })}\right| \le \bar{A}e^{-\bar{\omega } t},\quad s\ge 0, \quad y\in \mathbb {R}, \end{aligned}$$

for some \(\underline{A}>0\), \(\bar{A}>0\), \(\underline{\xi }\in \mathbb {R}\) and \(\bar{\xi }\in \mathbb {R}\). In particular, since \(\phi _{f}(-\infty )=1\) and \(\phi _{f}(+\infty )=0\), there exist \(T>0\) and \(B>0\) such that, for all \(s\ge T\),

$$\begin{aligned} \begin{aligned}&\underline{v}_{\beta }(s,y)>\alpha \quad \mathrm{if}~y\le c_{f}s-B,\\&\bar{v}_{\alpha }(s,y)<\beta \quad \mathrm{if}~y\ge c_{f}s+B. \end{aligned} \end{aligned}$$

It follows from (54) that for all \(t_{0}<t_{0}+T\le t\),

$$\begin{aligned} \begin{aligned}&u(t,x)>\alpha \quad \mathrm{if}~x\cdot e-\xi _{t_{0}}^{\epsilon }+M\le c_{f}(t-t_{0})-B,\\&u(t,x)<\beta \quad \mathrm{if}~x\cdot e-\xi _{t_{0}}^{\epsilon }-M\ge c_{f}(t-t_{0})+B. \end{aligned} \end{aligned}$$
(55)

By (45) and (55), for all \(t_{0}<t_{0}+T\le t\), we have

$$\begin{aligned} \begin{aligned}&\xi _{t_{0}}^{\epsilon }-M+c_{f}(t-t_{0})-B<\xi _{t}^{\epsilon }+M,\\&\xi _{t_{0}}^{\epsilon }+M+c_{f}(t-t_{0})+B>\xi _{t}^{\epsilon }-M. \end{aligned} \end{aligned}$$
(56)

By fixing \(t=0\), one gets that \(\limsup \limits _{t_{0}\rightarrow -\infty }\vert \xi _{t_{0}}^{\epsilon }-c_{f}t_{0}\vert \le \vert \xi _{0}^{\epsilon }\vert +2M+B\). For any arbitrary \(t\in \mathbb {R}\), letting \(t_{0}\rightarrow -\infty \) in (56) then leads to

$$\begin{aligned} \vert \xi _{t}^{\epsilon }-c_{f}t\vert \le \vert \xi _{0}\vert +4M+2B. \end{aligned}$$

Thus, by Definition 1 and (40), our solution \(u:\mathbb {R}\times \mathbb {R}^{N}\rightarrow (0,1)\) of (1) satisfies

$$\begin{aligned} \inf _{(t,x)\in \mathbb {R}\times \mathbb {R}^{N},x\cdot e-c_{f}t\le -A}u(t,x)\rightarrow 1~and~ \sup _{(t,x)\in \mathbb {R}\times \mathbb {R}^{N},x\cdot e-c_{f}t\ge A}u(t,x)\rightarrow 0\quad as~A\rightarrow +\infty . \end{aligned}$$

It follows from Theorem 3.1 of [3] and the uniqueness of the planar fronts that there exists \(\xi \in \mathbb {R}\) such that \(u(t,x)=\phi _{f}(x\cdot e-c_{f}t+\xi )\) for all \((t,x)\in \mathbb {R}\times \mathbb {R}^{N}\). This completes the proof of Theorem 2.

4 Existence of Non-standard Transition Fronts

In this section, we prove Theorem 3. That is, we prove the existence of new kinds of transition fronts, which are not invariant in any moving frame. We first consider the case \(N=2\) and construct two-dimensional transition fronts satisfying the conclusion of Theorem 3. The conclusion in \(\mathbb {R}^{N}\) with \(N>2\) will be then obtained immediately by trivially extending the constructed two-dimensional fronts in variables \(x_{3},\ldots ,x_{N}\). Now we first give some preliminaries.

For the standard planar traveling fronts \(\phi _{f}\), it is well known that there exist some positive constants \(\lambda _{1}\), \(C_{0}\), \(C_{1}\) and \(C_{2}\) such that

$$\begin{aligned} \phi _{f}(s)\le C_{0}e^{-c_{f}s},\quad s\ge 0, \end{aligned}$$
(57)
$$\begin{aligned} 1-\phi _{f}(s)\le C_{1}e^{\lambda _{1}s},\quad s\le 0, \end{aligned}$$
(58)
$$\begin{aligned} \vert \phi '_{f}(s)\vert \le C_{2}e^{-\lambda _{1}\vert s\vert },\quad s\in \mathbb {R}. \end{aligned}$$
(59)

Fix an angle \(\alpha \) such that \(\frac{\pi }{4}<\alpha <\frac{\pi }{2}\). Consider the quasilinear parabolic equation

$$\begin{aligned} W_{t}=\frac{W_{xx}}{1+W_{x}^{2}}+c_{f}\sqrt{1+W_{x}^{2}},\quad x\in \mathbb {R},\quad t>0. \end{aligned}$$
(60)

It follows from Propositions 1.1 and 2.5 of [27] that for any \(c>c_{f}\), there exists a unique solution \(\varphi (x;c)\) of (60) with asymptotic lines \(y=|x|\cot \alpha \) satisfying

$$\begin{aligned} c=\frac{\varphi _{xx}}{1+\varphi _{x}^{2}}+c_{f}\sqrt{1+\varphi _{x}^{2}},\quad x\in \mathbb {R}. \end{aligned}$$

Lemma 5

(Brazhnik [6], Ninomiya and Taniguchi [24, 26, 27]) There exist positive constants \(\gamma _{1}\), \(k_{i}\)\((i=1,2,3)\) and \(\omega _{\pm }\) such that

$$\begin{aligned}&\max \left\{ |\varphi ''(x)|,~|\varphi '''(x)|\right\} \le k_{1}\mathrm{sech}(\gamma _{1} x),\\&k_{2}\mathrm{sech}(\gamma _{1} x)\le \frac{c}{\sqrt{1+\varphi '(x)^{2}}}-c_{f}\le k_{3}\mathrm{sech}(\gamma _{1} x),\\&|x|\cot \alpha \le \varphi (x),~~~~~~~~~~~~~~~~~~\\&\omega _{-}\le \tilde{\omega }(x)\le \omega _{+}~~~~~~~~~~~~~~~~~ \end{aligned}$$

for any \(x\in \mathbb {R}\), where

$$\begin{aligned} \tilde{\omega }(x)=\frac{c(\varphi (x)-|x|\cot \alpha )}{c-c_{f}\sqrt{1+\varphi '(x)^{2}}}. \end{aligned}$$

By Lemma 5, it is easy to obtain that

$$\begin{aligned} 1>\frac{1}{\sqrt{1+\varphi '(x)^{2}}}>\frac{c_{f}}{c}\quad \mathrm{for~all}~x\in \mathbb {R} \end{aligned}$$

and that there exists a constant \(a>0\) such that

$$\begin{aligned} |x|\cot \alpha \le \varphi (x)\le |x|\cot \alpha +a\quad \mathrm{for~all}~x\in \mathbb {R}. \end{aligned}$$

It follows from [12, 36] that there exists a unique V-shaped traveling front \(\phi (x_{1},x_{2}-ct)\) (Fig. 1) of the problem (1) in \(\mathbb {R}^{2}\) satisfying the following properties: \(0<\phi <1\) in \(\mathbb {R}^{2}\), \(\phi \) is of class \(C^{2}(\mathbb {R}^{2})\), \(c=\frac{c_{f}}{\sin \alpha }\) and

$$\begin{aligned} \begin{aligned}&\liminf \limits _{A\rightarrow +\infty }\left( \inf \limits _{x_{2}\le \vert x_{1}\vert \cot \alpha -A}\phi (x_{1},x_{2})\right) =1,\\&\liminf \limits _{A\rightarrow +\infty }\left( \inf \limits _{x_{2}\ge \vert x_{1}\vert \cot \alpha +A}\phi (x_{1},x_{2})\right) =0. \end{aligned} \end{aligned}$$
(61)
Fig. 1
figure 1

The profiles (left figure) and the contour lines (right figure) of the V-shaped traveling front

Furthermore, for any \(\beta _{1}\in (0,1)\), there exist two positive constants \(\varepsilon _{0}^{+}(\beta _{1})\) and \(\alpha _{0}^{+}(\beta _{1},\varepsilon )\) so that, for \(0<\varepsilon <\varepsilon _{0}^{+}(\beta _{1})\) and \(0<\vartheta <\alpha _{0}^{+}(\beta _{1},\varepsilon )\),

$$\begin{aligned}&\phi _{f}\left( x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha \right) \nonumber \\&\quad<\phi (x_{1},x_{2})\nonumber \\&\quad < \phi _{f}\left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}\right) +\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\varphi (\vartheta x_{1})/\vartheta )\right) . \end{aligned}$$
(62)

Fix \(\beta _{1}\in (0,1)\), \(0<\varepsilon <\varepsilon _{0}^{+}(\beta _{1})\) and \(0<\vartheta <\alpha _{0}^{+}(\beta _{1},\varepsilon )\). Now we show that \(\phi \) is asymptotically planar along the directions \((\pm \sin \alpha ,\cos \alpha )\). This property plays an important role in the proof of Theorem 3.

Proposition 1

There exist two positive constants \(\rho _{1}\) and \(\omega _{1}\) such that

$$\begin{aligned} 0\le \phi (x_{1},x_{2})-\phi _{f}(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha )\le \rho _{1} e^{-\omega _{1}\sqrt{x_{1}^{2}+x_{2}^{2}}}\quad \mathrm{for~all}~(x_{1},x_{2})\in \mathbb {R}^{2}. \end{aligned}$$
(63)

Proof

Let

$$\begin{aligned} \rho _{1}=&\max \bigg \{C_{1},2\left( C_{2}\max _{s\in (-\infty ,0]} \left| se^{\lambda _{1}\frac{c_{f}}{c}s}\right| \frac{k_{3}}{c}+\max _{s\in \mathbb {R}} \left| \phi _{f}'(s)\right| \frac{\omega _{+}k_{3}}{c}+\varepsilon \right) ,\\&2\left( \frac{\omega _{+}k_{3}}{c}+ \beta _{1}\frac{c_{f}}{c}a\varepsilon \right) C_{2}e^{\lambda _{1}a}+2C_{0}^{\beta _{1}}\bigg \}. \end{aligned}$$

Choose \(\mu \in (0,1)\) such that

$$\begin{aligned} \mu \frac{\lambda _{1}c_{f}\cot \alpha }{c}<\frac{1}{2}\gamma _{1}\vartheta \quad \mathrm{and}\quad \mu \frac{\beta _{1}c_{f}^{2}\cot \alpha }{c}<\frac{1}{2}\gamma _{1}\vartheta . \end{aligned}$$

Fix a real number \(\omega _{1}\) such that

$$\begin{aligned} 0<\omega _{1}<\min \left\{ \frac{\lambda _{1}c_{f}\cot \alpha }{c},\frac{1}{2}\gamma _{1}\vartheta , \frac{\gamma _{1}\vartheta }{2\cot \alpha },\frac{\mu c_{f}\lambda _{1}}{c}, \frac{\mu \beta _{1}c_{f}^{2}}{c}\right\} . \end{aligned}$$

Now we divide our proof into three cases.

Case 1 when \(x_{2}\le 0\), by (58), one has

$$\begin{aligned} 0&<\phi (x_{1},x_{2})-\phi _{f}(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha )\\&<1-\phi _{f}(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha )\\&\le C_{1}e^{\lambda _{1}(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha )}\\&\le \rho _{1}e^{-\omega _{1}\sqrt{x_{1}^{2}+x_{2}^{2}}}. \end{aligned}$$

Case 2 when \(x_{2}>0\) and \(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha <0\), one has

$$\begin{aligned} x_{1}^{2}>\frac{x_{2}^{2}}{\cot ^{2}\alpha }~~\mathrm{and}~~\frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}-\frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )<0 \end{aligned}$$

from \(\frac{1}{\sqrt{1+\vert \varphi '(x_{1})\vert ^{2}}}>\frac{c_{f}}{c}\) and \(|x_{1}|\cot \alpha \le \varphi (x_{1})\) for any \(x_{1}\in \mathbb {R}\). Thus it follows from (59), Lemma 5 and \(c=\frac{c_{f}}{\sin \alpha }\) that

$$\begin{aligned} 0<\,&\phi (x_{1},x_{2})-\phi _{f}(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha )\\<\,&\phi _{f}\left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}}\right) +\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\varphi (\vartheta x_{1})/\vartheta )\right) \\&-\phi _{f}\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \\<&\int _{0}^{1}\phi _{f}^{\prime }\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )+\left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}}-\frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \upsilon \right) d\upsilon \\&\times \left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}}-\frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) +\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\\ =&\int _{0}^{1}\phi _{f}^{\prime }\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )+\left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}}-\frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \upsilon \right) d\upsilon \\&\times \left( \frac{1}{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}}-\frac{c_{f}}{c}\right) (x_{2}-\vert x_{1}\vert \cot \alpha )\\&-\int _{0}^{1}\phi _{f}^{\prime }\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )+\left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}}-\frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \upsilon \right) d\upsilon \\&\times \frac{\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha }{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}} +\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\\ \le&-C_{2}e^{\lambda _{1}\frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )}\left( \frac{1}{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}}-\frac{c_{f}}{c}\right) (x_{2}-\vert x_{1}\vert \cot \alpha )\\&+\max _{s\in \mathbb {R}} \left| \phi _{f}^{\prime }(s)\right| \frac{\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha }{\sqrt{1+\vert \varphi ^{\prime }(\vartheta x_{1})\vert ^{2}}} +\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\\ \le&\left( C_{2}\max _{s\in (-\infty ,0]} \left| se^{\lambda _{1}\frac{c_{f}}{c}s}\right| \frac{k_{3}}{c}+\max _{s\in \mathbb {R}} \left| \phi _{f}^{\prime }(s)\right| \frac{\omega _{+}k_{3}}{c}+\varepsilon \right) \mathrm{sech}(\gamma _{1}\vartheta x_{1})\\ <\,&2\left( C_{2}\max _{s\in (-\infty ,0]}\left| se^{\lambda _{1}\frac{c_{f}}{c}s}\right| \frac{k_{3}}{c}+\max _{s\in \mathbb {R}} \left| \phi _{f}^{\prime }(s)\right| \frac{\omega _{+}k_{3}}{c}+\varepsilon \right) e^{-\gamma _{1}\vartheta \vert x_{1}\vert }\\ \le \,&\rho _{1}e^{-\omega _{1}\sqrt{x_{1}^{2}+x_{2}^{2}}}. \end{aligned}$$

Case 3 when \(x_{2}>0\) and \(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha \ge 0\), one has \(x_{2}>\vert x_{1}\vert \cot \alpha \). It follows from (57), (59), Lemma 5 and \(c=\frac{c_{f}}{\sin \alpha }\) that

$$\begin{aligned} 0<\,&\phi (x_{1},x_{2})-\phi _{f}(x_{2}\sin \alpha -\vert x_{1}\vert \cos \alpha )\\<\,&\phi _{f}\left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}\right) +\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\varphi (\vartheta x_{1})/\vartheta )\right) \\&-\phi _{f}\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \\ \le&\phi _{f}\left( \frac{x_{2}-\varphi (\vartheta x_{1})/\vartheta }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}\right) -\phi _{f}\left( \frac{x_{2}-\vert x_{1}\vert \cot \alpha }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}\right) \\&+\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\varphi (\vartheta x_{1})/\vartheta )\right) \\ =&\int _{0}^{1}-\phi _{f}'\left( \frac{x_{2}-\vert x_{1}\vert \cot \alpha -(\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha )\upsilon }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}\right) d\upsilon \\&\times \frac{\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}+\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\varphi (\vartheta x_{1})/\vartheta )\right) \\&-\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \\&+\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \\ =&\int _{0}^{1}-\phi _{f}'\left( \frac{x_{2}-\vert x_{1}\vert \cot \alpha -(\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha )\upsilon }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}\right) d\upsilon \\&\times \frac{\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}\\&+\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\int _{0}^{1}-\beta _{1}\phi _{f}'\left( \frac{c_{f}}{c}\left( x_{2}-\vert x_{1}\vert \cot \alpha -(\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha )\upsilon \right) \right) \\&\times \phi _{f}^{\beta _{1}-1}\left( \frac{c_{f}}{c}\left( x_{2}-\vert x_{1}\vert \cot \alpha -(\varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha )\upsilon \right) \right) d\upsilon \\&\times \frac{c_{f}}{c}\left( \varphi (\vartheta x_{1})/\vartheta -\vert x_{1}\vert \cot \alpha \right) +\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})\phi _{f}^{\beta _{1}}\left( \frac{c_{f}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )\right) \\ \le \,&C_{2}e^{-\lambda _{1}\frac{x_{2}-\vert x_{1}\vert \cot \alpha }{\sqrt{1+\vert \varphi '(\vartheta x_{1})\vert ^{2}}}}e^{\lambda _{1}a}\frac{\omega _{+}k_{3}}{c}\mathrm{sech}(\gamma _{1}\vartheta x_{1})\\&+\beta _{1}C_{2}e^{-\lambda _{1}\frac{c_{f}}{c}\left( x_{2}-\vert x_{1}\vert \cot \alpha \right) }e^{\lambda _{1}a}\frac{c_{f}}{c}a\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1}) \\&+\varepsilon \mathrm{sech}(\gamma _{1}\vartheta x_{1})C_{0}^{\beta _{1}}e^{-\beta _{1}\frac{c_{f}^{2}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )}\\ \le \,&2\left( \frac{\omega _{+}k_{3}}{c}+\beta _{1} \frac{c_{f}}{c}a\varepsilon \right) C_{2}e^{\lambda _{1}a}e^{-\frac{c_{f}}{c}\lambda _{1}(x_{2} -\vert x_{1}\vert \cot \alpha )}e^{-\gamma _{1}\vartheta \vert x_{1}\vert }\\&+2C_{0}^{\beta _{1}} e^{-\frac{\beta _{1}c_{f}^{2}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )}e^{-\gamma _{1}\vartheta \vert x_{1}\vert }\\ \le \,&2\left( \frac{\omega _{+}k_{3}}{c}+\beta _{1} \frac{c_{f}}{c}a\varepsilon \right) C_{2}e^{\lambda _{1}a}e^{-\mu \frac{c_{f}}{c} \lambda _{1}(x_{2}-\vert x_{1}\vert \cot \alpha )}e^{-\gamma _{1}\vartheta \vert x_{1}\vert }\\&+2C_{0}^{\beta _{1}}e^{-\mu \frac{\beta _{1}c_{f}^{2}}{c}(x_{2}-\vert x_{1}\vert \cot \alpha )}e^{-\gamma _{1}\vartheta \vert x_{1}\vert }\\ \le \,&2\left( \frac{\omega _{+}k_{3}}{c}+\beta _{1} \frac{c_{f}}{c}a\varepsilon \right) C_{2}e^{\lambda _{1}a}e^{-\mu \frac{c_{f}}{c}\lambda _{1}x_{2}-\frac{1}{2}\gamma _{1}\vartheta \vert x_{1}\vert }+2C_{0}^{\beta _{1}}e^{-\mu \frac{\beta _{1}c_{f}^{2}}{c}x_{2}-\frac{1}{2}\gamma _{1}\vartheta \vert x_{1}\vert }\\ <\,&\left[ 2\left( \frac{\omega _{+}k_{3}}{c}+\beta _{1} \frac{c_{f}}{c}a\varepsilon \right) C_{2}e^{\lambda _{1}a}+2C_{0}^{\beta _{1}}\right] e^{-\omega _{1}(\vert x_{1}\vert +\vert x_{2}\vert )}\\ \le \,&\rho _{1}e^{-\omega _{1}\sqrt{x_{1}^{2}+x_{2}^{2}}}. \end{aligned}$$

Combining the above three cases, the proof of Proposition 1 is thereby complete. \(\square \)

It follows from Proposition 1 and the Schauder interior estimates that there exist two positive constants \(\rho _{2}\) and \(\omega _{2}\) such that

$$\begin{aligned} \left| \nabla \phi (x_{1},x_{2})-\nabla (\phi _{f}(x_{2}\sin \alpha -x_{1}\cos \alpha ))\right| \le \rho _{2}e^{-\omega _{2}\sqrt{x_{1}^{2}+x_{2}^{2}}}\quad \mathrm{for~all}\quad x_{1}\ge 0,x_{2}\in \mathbb {R}. \end{aligned}$$

Whence

$$\begin{aligned} \begin{aligned}&\left| \phi _{x_{1}}(x_{1},x_{2})+\phi _{f}^{\prime }(x_{2}\sin \alpha -x_{1}\cos \alpha )\cos \alpha \right| \le \rho _{2}e^{-\omega _{2}\sqrt{x_{1}^{2}+x_{2}^{2}}},\\&\left| \phi _{x_{2}}(x_{1},x_{2})-\phi _{f}^{\prime }(x_{2}\sin \alpha -x_{1}\cos \alpha )\sin \alpha \right| \le \rho _{2}e^{-\omega _{2}\sqrt{x_{1}^{2}+x_{2}^{2}}} \end{aligned} \end{aligned}$$
(64)

for all \(x_{1}\ge 0,~x_{2}\in \mathbb {R}\). Since the standard planar traveling fronts \(\phi _{f}(s)\) converges exponentially fast to 0 and 1 as \(s\rightarrow \pm \infty \), Proposition 1 yields that the V-shaped traveling front \(\phi \) also converges exponentially fast to 0 and 1 as \(x_{2}-\vert x_{1}\vert \cot \alpha \rightarrow \pm \infty \). By the Schauder interior estimates, there exist two positive constants \(\rho _{3}\) and \(\omega _{3}\) such that

$$\begin{aligned} \vert \nabla \phi (x_{1},x_{2})\vert \le \rho _{3} e^{-\omega _{3}\vert x_{2}-\vert x_{1}\vert \cot \alpha \vert }\quad \mathrm{for~all~}(x_{1},x_{2})\in \mathbb {R}^{2}. \end{aligned}$$
(65)

It follows from Corollary 3.3 and Lemma 3.4 of [36] that

$$\begin{aligned} \forall ~A\ge 0,\quad \sup _{-A\le x_{2}-\vert x_{1}\vert \cot \alpha \le A}\phi _{x_{2}}(x_{1},x_{2})<0 \end{aligned}$$
(66)

and that \(\phi \) is decreasing in any direction \((\cos \hat{\alpha },\sin \hat{\alpha })\) such that \(\pi /2-\alpha<\hat{\alpha }<\pi /2+\alpha \), see also [12]. In particular, the function \(\phi \) is nonincreasing along the directions \((\pm \sin \alpha ,\cos \alpha )\).

Fig. 2
figure 2

The profiles (left figure) and the contour lines (right figure) of function \(\psi \)

Define

$$\begin{aligned} \psi (x_{1},x_{2})=\phi (x_{1}\sin \alpha -x_{2}\cos \alpha ,x_{1}\cos \alpha +x_{2}\sin \alpha )\quad \mathrm{for~all}~(x_{1},x_{2})\in \mathbb {R}^{2}, \end{aligned}$$
(67)

which rotates the function \(\phi \) with angle \(\alpha -\frac{\pi }{2}\) clockwise. Then the function \(\psi \) (Fig. 2) is decreasing in any direction \((\cos \hat{\beta },\sin \hat{\beta })\) with \(0<\hat{\beta }<2\alpha \). In particular, \(\psi \) is nonincreasing in the horizontal direction (1, 0) and it converges to the planar front \(\phi _{f}(x_{2})\) along this direction. Set

$$\begin{aligned} \underline{v}(t,x_{1},x_{2})=\,&\psi (x_{1}-ct\cos \alpha ,x_{2}-ct\sin \alpha )\nonumber \\ =\,&\phi (x_{1}\sin \alpha -x_{2}\cos \alpha ,x_{1}\cos \alpha +x_{2}\sin \alpha -ct). \end{aligned}$$
(68)

Since \(\phi (x_{1},x_{2}-ct)\) solves the Eq. (1) in \(\mathbb {R}^{2}\), then the \(C^{2}(\mathbb {R}\times \mathbb {R}^{2})\) function \(\underline{v}\) also satisfies (1) in \(\mathbb {R}^{2}\). Moreover, the definition of \(\underline{v}\) yields \(\underline{v}_{t}(t,x_{1},x_{2})>0\) and \(\underline{v}_{x_{1}}(t,x_{1},x_{2})\le 0\) in \(\mathbb {R}\times \mathbb {R}^{2}\).

Now we consider the following Neumann boundary value problem in half-space \(H=\{(x_{1},x_{2})\in \mathbb {R}^{2},x_{1}<0\}\)

$$\begin{aligned} \begin{aligned} v_{t}&=\varDelta v+f(v), (t,x_{1},x_{2})\in \mathbb {R}\times H,\\ v_{x_{1}}&=0, (t,x_{1},x_{2})=(t,0,x_{2})\in \mathbb {R}\times \partial H. \end{aligned} \end{aligned}$$
(69)

It is easy to see that the function \(\underline{v}\) is a subsolution of (69).

In the following lemma, we construct a supersolution which looks like the function \(\underline{v}\) for very negative times, up to some exponentially small terms.

Lemma 6

There exist some constants \(\sigma >0\), \(\delta >0\) and \(T<0\) such that the function

$$\begin{aligned} \bar{v}(t,x_{1},x_{2})=\min \left\{ \underline{v}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right) +\delta e^{\delta (x_{1}+t)},1\right\} \end{aligned}$$
(70)

is a supersolution of (69) for \(t\le T\) and \((x_{1},x_{2})\in \overline{H}\).

Proof

Let

$$\begin{aligned} \omega _{4}=\frac{\min (\omega _{2}c_{f}\cos \alpha ,\omega _{3}c)}{2}>0, \end{aligned}$$

where \(\omega _{2}\) and \(\omega _{3}\) are given in (64) and (65). Choose \(\delta \) such that

$$\begin{aligned} 0<\delta <\min (1,\omega _{4},\theta /2)~~\mathrm{and}~~f'\le 0~~\mathrm{on}~[1-\delta ,1]. \end{aligned}$$
(71)

It follows from (61) that there exists a real number \(A>0\) such that

$$\begin{aligned} \begin{array}{ll} \phi (x_{1},x_{2})\ge 1-\delta &{}\quad \mathrm{for~all} ~x_{2}\le \vert x_{1}\vert \cot \alpha -A,\\ \phi (x_{1},x_{2})\le \delta &{}\quad \mathrm{for~all} ~x_{2}\ge \vert x_{1}\vert \cot \alpha +A. \end{array} \end{aligned}$$
(72)

Equation (66) implies that there exists a constant \(\kappa >0\) such that

$$\begin{aligned} \sup _{-A\le x_{2}-\vert x_{1}\vert \cot \alpha \le A}\phi _{x_{2}}(x_{1},x_{2})=-\kappa <0. \end{aligned}$$
(73)

Choose \(\sigma >0\) such that

$$\begin{aligned} \sigma c\kappa \ge L=\max _{u\in [0,1]}\vert f'(u)\vert . \end{aligned}$$
(74)

Set

$$\begin{aligned} \rho _{4}=(\sin \alpha +\cos \alpha )\max (\rho _{2},\rho _{3}e^{\omega _{3}c\sigma })>0. \end{aligned}$$

Let \(T<0\) be such that

$$\begin{aligned} T\le -2\sigma <0~~\mathrm{and}~~\delta ^{2}e^{\delta t}\ge \rho _{4}e^{\omega _{4}t}\quad \mathrm{for~all}~t\le T. \end{aligned}$$

Similar to the Lemma 5.1 of Hamel [11] combining with Proposition 1, we can prove that \(\bar{v}_{x_{1}}\ge 0\) on \((-\infty ,T]\times \partial H\) in the region where \(\bar{v}<1\).

Since \(f(1)=0\), it is sufficient to show that

$$\begin{aligned} \bar{v}_{t}\ge \varDelta \bar{v}+f(\bar{v}) \end{aligned}$$

on the region \((t,x_{1},x_{2})\in (-\infty ,T]\times \overline{H}\) such that \(\bar{v}<1\). Since \(\underline{v}\) satisfies (1) in \(\mathbb {R}^{2}\) and \(\delta <1\), thus

$$\begin{aligned} \overline{\mathscr {L}}(t,x_{1},x_{2}):=\,&\bar{v}_{t}(t,x_{1},x_{2}) -\varDelta \bar{v}(t,x_{1},x_{2})-f(\bar{v}(t,x_{1},x_{2}))\nonumber \\ =\,&\underline{v}_{t}(t+\sigma e^{\delta t},x_{1},x_{2})+\sigma \delta \underline{v}_{t}(t+\sigma e^{\delta t},x_{1},x_{2})e^{\delta t}+\delta ^{2}e^{\delta (x_{1}+t)}\nonumber \\&-\varDelta \underline{v}(t+\sigma e^{\delta t},x_{1},x_{2})-\delta ^{3}e^{\delta (x_{1}+t)}-f\left( \bar{v}(t,x_{1},x_{2})\right) \nonumber \\ \ge \,&f\left( \underline{v}(t+\sigma e^{\delta t},x_{1},x_{2})\right) -f\left( \bar{v}(t,x_{1},x_{2})\right) +\sigma \delta \underline{v}_{t}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right) e^{\delta t} \end{aligned}$$
(75)

For simplicity, by (68), we can set

$$\begin{aligned} \underline{v}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right) =\phi (\xi _{1}(x_{1},x_{2}),\xi _{2}(t,x_{1},x_{2})), \end{aligned}$$

where

$$\begin{aligned} \xi _{1}(x_{1},x_{2})=x_{1}\sin \alpha -x_{2}\cos \alpha ~~\mathrm{and}~~ \xi _{2}(t,x_{1},x_{2})=x_{1}\cos \alpha +x_{2}\sin \alpha -ct-c\sigma e^{\delta t}. \end{aligned}$$

Firstly, if \(\xi _{2}(t,x_{1},x_{2})\le \vert \xi _{1}(x_{1},x_{2})\vert \cot \alpha -A\), then (72) implies that

$$\begin{aligned} 1>\bar{v}(t,x_{1},x_{2})>\underline{v}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right) =\phi (\xi _{1}(x_{1},x_{2}),\xi _{2}(t,x_{1},x_{2}))\ge 1-\delta . \end{aligned}$$

It follows from (71), (75) and \(\underline{v}_{t}>0\) that one has

$$\begin{aligned} \overline{\mathscr {L}}(t,x_{1},x_{2})&\ge f\left( \underline{v}(t+\sigma e^{\delta t},x_{1},x_{2})\right) -f\left( \bar{v}(t,x_{1},x_{2})\right) +\sigma \delta \underline{v}_{t}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right) e^{\delta t}\nonumber \\&\ge 0. \end{aligned}$$
(76)

Secondly, if \(\xi _{2}(t,x_{1},x_{2})\ge \vert \xi _{1}(x_{1},x_{2})\vert \cot \alpha +A\), then it follows from (71), (72), \(x_{1}\le 0\) and \(t\le T<0\) that

$$\begin{aligned} 0<\underline{v}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right)<\bar{v}(t,x_{1},x_{2})=\underline{v}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right) +\delta e^{\delta (x_{1}+t)}\le 2\delta <\theta . \end{aligned}$$

Since \(f=0\) on \([0,\theta ]\) and \(\underline{v}_{t}>0\), then

$$\begin{aligned} \overline{\mathscr {L}}(t,x_{1},x_{2})&\ge f\left( \underline{v}(t+\sigma e^{\delta t},x_{1},x_{2})\right) -f\left( \bar{v}(t,x_{1},x_{2})\right) +\sigma \delta \underline{v}_{t}\left( t+\sigma e^{\delta t},x_{1},x_{2}\right) e^{\delta t}\nonumber \\&\ge 0. \end{aligned}$$
(77)

Lastly, if \(-A\le \xi _{2}(t,x_{1},x_{2})-\vert \xi _{1}(x_{1},x_{2})\vert \cot \alpha \le A\), then

$$\begin{aligned}&f\left( \underline{v}(t+\sigma e^{\delta t},x_{1},x_{2})\right) -f\left( \bar{v}(t,x_{1},x_{2})\right) \\&\quad =f\left( \underline{v}(t+\sigma e^{\delta t},x_{1},x_{2})\right) -f\left( \underline{v}(t+\sigma e^{\delta t},x_{1},x_{2})+\delta e^{\delta (x_{1}+t)}\right) \\&\quad \ge -L\delta e^{\delta (x_{1}+t)} \end{aligned}$$

and

$$\begin{aligned} \underline{v}_{t}(t+\sigma e^{\delta t},x_{1},x_{2})=-c\phi _{x_{2}}(\xi _{1}(x_{1},x_{2}),\xi _{2}(t,x_{1},x_{2}))\ge c\kappa . \end{aligned}$$

It follows from (74) and \(x_{1}\le 0\) that

$$\begin{aligned} \overline{\mathscr {L}}(t,x_{1},x_{2})\ge -L\delta e^{\delta (x_{1}+t)}+\sigma \delta c\kappa e^{\delta t}\ge \delta (\sigma c\kappa -L)e^{\delta t}\ge 0. \end{aligned}$$
(78)

Combining with (76), (77) and (78), one has \(\overline{\mathscr {L}}(t,x_{1},x_{2})\ge 0\) for all \((t,x_{1},x_{2})\in (-\infty ,T]\times \overline{H}\) such that \(\bar{v}(t,x_{1},x_{2})<1\). This completes the proof. \(\square \)

Proof of Theorem 3

It follows from the positivity of \(\underline{v}_{t}\) and the definition of \(\bar{v}\) that \(\underline{v}(t,x_{1},x_{2})<\bar{v}(t,x_{1},x_{2})\) in \(\mathbb {R}\times \overline{H}\). For any \(n\in \mathbb {N}\) such that \(n>\vert T\vert \), let \(v^{n}\) be the solution of the Cauchy problem associated to (69) for times \(t>-n\), with initial condition

$$\begin{aligned} v^{n}(-n,x_{1},x_{2})=\underline{v}(-n,x_{1},x_{2})\quad \mathrm{for~all}~(x_{1},x_{2})\in H. \end{aligned}$$

Since \((\underline{v},\bar{v})\) is a couple of sub-supersolution of the problem (69), the maximum principle implies that

$$\begin{aligned} 0<\underline{v}(t,x_{1},x_{2})\le v^{n}(t,x_{1},x_{2})\le \bar{v}(t,x_{1},x_{2})\le 1 \end{aligned}$$

for all \(-n<t\le T\) and \((x_{1},x_{2})\in \overline{H}\) and that

$$\begin{aligned} 0<\underline{v}(t,x_{1},x_{2})\le v^{n}(t,x_{1},x_{2})\le 1~~\mathrm{for~all~}(t,x_{1},x_{2})\in (-n,+\infty )\times \overline{H}. \end{aligned}$$
(79)

In particular, for every \((t,x_{1},x_{2})\in \mathbb {R}\times \overline{H}\), the sequence \((v^{n}(t,x_{1},x_{2}))_{n>\max (\vert T\vert ,\vert t\vert )}\) is nondecreasing. Furthermore, since \(\underline{v}_{t}>0\), (79) and the maximum principle yield that \(v^{n}\) is increasing with respect to time t in \(\overline{H}\).

It follows from monotone convergence and standard parabolic estimates up to the boundary that the functions \(v^{n}\) converge to a solution v of (69) as \(n\rightarrow +\infty \) in \(C_{loc}^{1,2}(\mathbb {R}\times \overline{H})\). Furthermore, one has

$$\begin{aligned} 0<\underline{v}(t,x_{1},x_{2})\le v(t,x_{1},x_{2})\le \bar{v}(t,x_{1},x_{2})\le 1\quad \mathrm{for~all~}t\le T~\mathrm{and}~(x_{1},x_{2})\in \overline{H}, \end{aligned}$$
(80)

and

$$\begin{aligned} 0<\underline{v}\le v\le 1,~~v_{t}\ge 0\quad \mathrm{in}~\mathbb {R}\times \overline{H}. \end{aligned}$$

In particular, since for each fixed \((x_{1},x_{2})\in \overline{H}\), the function \(\bar{v}(t,x_{1},x_{2})\rightarrow 0<1\) as \(t\rightarrow -\infty \), then it follows from (80) and the strong maximum principle that \(0<v<1\) in \(\mathbb {R}\times \overline{H}\).

Now we construct a solution u of (1) in \(\mathbb {R}^{2}\). Define u in \(\mathbb {R}\times \mathbb {R}^{2}\) as

$$\begin{aligned} u(t,x_{1},x_{2})= {\left\{ \begin{array}{ll} v(t,x_{1},x_{2})\quad t\in \mathbb {R},\quad x_{1}\le 0,\quad x_{2}\in \mathbb {R},\\ v(t,-x_{1},x_{2})\quad t\in \mathbb {R},\quad x_{1}>0,\quad x_{2}\in \mathbb {R}. \end{array}\right. } \end{aligned}$$

Since v satisfies (69) in the half-plane H with Neumann boundary conditions, then u is a classical time-global solution of (1) in the whole plane \(\mathbb {R}^{2}\). Furthermore, \(0<u<1\) in \(\mathbb {R}\times \mathbb {R}^{2}\),

$$\begin{aligned} \underline{v}(t,-\vert x_{1}\vert ,x_{2})\le u(t,x_{1},x_{2})\quad \mathrm{for~all~}(t,x_{1},x_{2})\in \mathbb {R}\times \mathbb {R}^{2} \end{aligned}$$

and

$$\begin{aligned} \underline{v}(t,-\vert x_{1}\vert ,x_{2})\le u(t,x_{1},x_{2})\le \bar{v}(t,-\vert x_{1}\vert ,x_{2})\quad \mathrm{for~all~}t\le T~\mathrm{and~}(x_{1},x_{2})\in \mathbb {R}^{2}. \end{aligned}$$

Therefore, by the definition of \(\underline{v}\) and (62), one has

$$\begin{aligned} \max (\phi _{f}(-\vert x_{1}\vert \sin (2\alpha )-x_{2}\cos (2\alpha )-c_{f}t),\phi _{f}(x_{2}-c_{f}t))\le u(t,x_{1},x_{2}) \end{aligned}$$

for all \((t,x_{1},x_{2})\in \mathbb {R}\times \mathbb {R}^{2}\). And it follows from the definition \(\bar{v}\) and Proposition 1 that

$$\begin{aligned}&u(t,x_{1},x_{2})\\&\quad \le \max (\phi _{f}(-\vert x_{1}\vert \sin (2\alpha )-x_{2}\cos (2\alpha )-c_{f}t-c_{f}\sigma e^{\delta t}),\phi _{f}(x_{2}-c_{f}t-c_{f}\sigma e^{\delta t}))\nonumber \\&\qquad +\rho _{1}e^{-\omega _{1}\sqrt{(\vert x_{1}\vert \sin \alpha +x_{2}\cos \alpha )^{2}+(\vert x_{1}\vert \cos \alpha -x_{2}\sin \alpha +ct+c\sigma e^{\delta t})^{2}}}+\delta e^{\delta (t-\vert x_{1}\vert )} \end{aligned}$$

for all \(t\le T\) and \((x_{1},x_{2})\in \mathbb {R}^{2}\).

For \(t\le 0\), let

$$\begin{aligned} \begin{aligned}&P_{t}^{l}=(ct\cos \alpha ,ct\sin \alpha ),\quad L_{t}^{l}=P_{t}^{l}+\mathbb {R}_{+}(\cos (2\alpha ),\sin (2\alpha )),\\&P_{t}^{r}=(-ct\cos \alpha ,ct\sin \alpha ),\quad L_{t}^{r}=P_{t}^{r}+\mathbb {R}_{+}(-\cos (2\alpha ),\sin (2\alpha )) \end{aligned} \end{aligned}$$

and

$$\begin{aligned} \varGamma _{t}=L_{t}^{l}\cup [P_{t}^{l},P_{t}^{r}]\cup L_{t}^{r}\quad \mathrm{for~all~}t\le 0, \end{aligned}$$
(81)

where the superscript l (resp. r) stands for left (resp. right). Define

$$\begin{aligned} \varGamma _{t}=\left\{ (x_{1},x_{2})\in \mathbb {R}^{2}\big \vert x_{2}=\vert \tan (2\alpha )\vert \vert x_{1}\vert +\frac{c_{f}t}{\vert \cos (2\alpha )\vert }\right\} \quad \mathrm{for~all~}t>0. \end{aligned}$$
(82)

Thus, for every \(t\in \mathbb {R}\), \(\varGamma _{t}\) can be written as a graph \(\varGamma _{t}=\{(x_{1},x_{2})\in \mathbb {R}^{2};x_{2}=\hat{\varphi }_{t}(x_{1})\}\), where \(\hat{\varphi }_{t}:\mathbb {R}\rightarrow \mathbb {R}\) is a Lipschitz-continuous function. For all \(t\in \mathbb {R}\), define

$$\begin{aligned} \varOmega _{t}^{+}=\{(x_{1},x_{2})\in \mathbb {R}^{2}\big \vert x_{2}<\hat{\varphi }_{t}(x_{1})\}~\mathrm{and}~ \varOmega _{t}^{-}=\{(x_{1},x_{2})\in \mathbb {R}^{2}\big \vert x_{2}>\hat{\varphi }_{t}(x_{1})\}. \end{aligned}$$
(83)

Obviously, the sets \((\varOmega _{t}^{\pm })_{t\in \mathbb {R}}\) and \((\varGamma _{t})_{t\in \mathbb {R}}\) satisfy the general properties (5) and (6).

Similar to the proof of Lemma 5.2 of [11], the function u is a transition front connecting 0 and 1 for problem (1) in \(\mathbb {R}^{2}\) with the sets \((\varOmega _{t}^{\pm })_{t\in \mathbb {R}}\) and \((\varGamma _{t})_{t\in \mathbb {R}}\).

Now we prove that the solution u is not invariant as time runs with any moving frame. That is, it satisfies the conclusion of Theorem 3. Assume by contradiction that there exist a function \(\varPhi :\mathbb {R}^{2}\rightarrow (0,1)\) and some families \((R_{t})_{t\in \mathbb {R}}\) and \((X_{t})_{t\in \mathbb {R}}=(x_{1,t},x_{2,t})_{t\in \mathbb {R}}\) of rotations and points in \(\mathbb {R}^{2}\) such that

$$\begin{aligned} u(t,x_{1},x_{2})=\varPhi (R_{t}(x_{1}-x_{1,t},x_{2}-x_{2,t}))\quad \mathrm{for~all~}(t,x_{1},x_{2})\in \mathbb {R}\times \mathbb {R}^{2}. \end{aligned}$$

Then there is \(M\ge 0\) such that

$$\begin{aligned} R_{t}(\varGamma _{t}-X_{t})\subset \{(x_{1},x_{2})\in \mathbb {R}^{2}\big \vert d((x_{1},x_{2}), R_{s}(\varGamma _{t}-X_{s}))\le M\}~\mathrm{for~all~}(t,s)\in \mathbb {R}^{2}, \end{aligned}$$

which is contradicted with the definitions of the sets \(\varGamma _{t}\) defined as (81) and (82). Whence, Theorem 3 holds in \(\mathbb {R}^{2}\).

Now, we extend the transition front u trivially in \(\mathbb {R}^{N}\)\((N\ge 3)\). Let

$$\begin{aligned} \widetilde{u}(t,x_{1},\ldots ,x_{N})=u(t,x_{1},x_{2})\quad \mathrm{for~all~}(t,x_{1},\ldots ,x_{N})\in \mathbb {R}\times \mathbb {R}^{N}. \end{aligned}$$

Obviously, the function \(\widetilde{u}\) is a transition front connecting 0 and 1 for problem (1) in \(\mathbb {R}^{N}\) with the sets

$$\begin{aligned} \widetilde{\varOmega }_{t}^{\pm }=\{(x_{1},\ldots ,x_{N})\in \mathbb {R}^{N}\big \vert (x_{1},x_{2})\in \varOmega _{t}^{\pm }\}\quad \mathrm{for~all}~t\in \mathbb {R} \end{aligned}$$

and satisfies the desired conclusion. This completes the proof of Theorem 3.