Loading [Contrib]/a11y/accessibility-menu.js
Research article

Barycentric interpolation collocation method based on Crank-Nicolson scheme for the Allen-Cahn equation

  • Received: 08 December 2020 Accepted: 18 January 2021 Published: 29 January 2021
  • MSC : 65M70, 65L20

  • This paper proposes a numerical scheme for the Allen-Cahn equation that represents a phenomenological model for anti-phase domain coarsening in a binary mixture. In order to obtain a high order discretization in space, we adopt the barycentric interpolation collocation method. The semi-discretized scheme in space is shown to be consistent. The second-order Crank-Nicolson scheme is used for temporal discretization and the simple iteration method is adopted for nonlinear term. Corresponding algebraic system is derived. Numerical examples are demonstrated to validate the efficiency of the proposed method.

    Citation: Yangfang Deng, Zhifeng Weng. Barycentric interpolation collocation method based on Crank-Nicolson scheme for the Allen-Cahn equation[J]. AIMS Mathematics, 2021, 6(4): 3857-3873. doi: 10.3934/math.2021229

    Related Papers:

    [1] Mengting Sui, Yanfei Du . Bifurcations, stability switches and chaos in a diffusive predator-prey model with fear response delay. Electronic Research Archive, 2023, 31(9): 5124-5150. doi: 10.3934/era.2023262
    [2] Yang Jin, Wanxiang Li, Hongbing Zhang . Transition characteristics of the dynamic behavior of a vehicle wheel-rail vibro-impact system. Electronic Research Archive, 2023, 31(11): 7040-7060. doi: 10.3934/era.2023357
    [3] Qinghua Zhu, Meng Li, Fang Han . Hopf bifurcation control of the ML neuron model with Hc bifurcation type. Electronic Research Archive, 2022, 30(2): 615-632. doi: 10.3934/era.2022032
    [4] Yuan Xue, Jinli Xu, Yuting Ding . Dynamics analysis of a diffusional immunosuppressive infection model with Beddington-DeAngelis functional response. Electronic Research Archive, 2023, 31(10): 6071-6088. doi: 10.3934/era.2023309
    [5] Qixiang Wen, Shenquan Liu, Bo Lu . Firing patterns and bifurcation analysis of neurons under electromagnetic induction. Electronic Research Archive, 2021, 29(5): 3205-3226. doi: 10.3934/era.2021034
    [6] San-Xing Wu, Xin-You Meng . Hopf bifurcation analysis of a multiple delays stage-structure predator-prey model with refuge and cooperation. Electronic Research Archive, 2025, 33(2): 995-1036. doi: 10.3934/era.2025045
    [7] Fengrong Zhang, Ruining Chen . Spatiotemporal patterns of a delayed diffusive prey-predator model with prey-taxis. Electronic Research Archive, 2024, 32(7): 4723-4740. doi: 10.3934/era.2024215
    [8] Mengxin He, Zhong Li . Dynamic behaviors of a Leslie-Gower predator-prey model with Smith growth and constant-yield harvesting. Electronic Research Archive, 2024, 32(11): 6424-6442. doi: 10.3934/era.2024299
    [9] Wenxin Zhang, Lijun Pei . Bifurcation and chaos in N-type and S-type muscular blood vessel models. Electronic Research Archive, 2025, 33(3): 1285-1305. doi: 10.3934/era.2025057
    [10] Weiyu Li, Hongyan Wang . Dynamics of a three-molecule autocatalytic Schnakenberg model with cross-diffusion: Turing patterns of spatially homogeneous Hopf bifurcating periodic solutions. Electronic Research Archive, 2023, 31(7): 4139-4154. doi: 10.3934/era.2023211
  • This paper proposes a numerical scheme for the Allen-Cahn equation that represents a phenomenological model for anti-phase domain coarsening in a binary mixture. In order to obtain a high order discretization in space, we adopt the barycentric interpolation collocation method. The semi-discretized scheme in space is shown to be consistent. The second-order Crank-Nicolson scheme is used for temporal discretization and the simple iteration method is adopted for nonlinear term. Corresponding algebraic system is derived. Numerical examples are demonstrated to validate the efficiency of the proposed method.



    The theory of bifurcations is crucial in understanding qualitative properties of differential systems depending on one or more parameters. Bifurcations of codimension one and two are well-addressed in the literature [1,2,3] when they are non-degenerate, but many open problems emerge when they become degenerate [4] or the codimension is beyond 2.

    The classical normal form of the double-Hopf bifurcation in differential systems of dimension 4 is based on six generic conditions (HH.0)–(HH.5) [1]. Some of the pioneering papers reporting results on non-degenerate double-Hopf bifurcation are references [5,6,7,8].

    If one or more generic conditions fail to be satisfied, a degeneracy arises and, thus, the generic results are not valid anymore. In this work we study two degenerate double-Hopf bifurcations. The degeneracy we refer at is related to generic conditions which are necessary in obtaining normal forms.

    Another added value of our results is that we discovered new bifurcation diagrams and new phase portraits and, thus, new properties of double-Hopf bifurcation which, to our knowledge, were not previously reported in the literature. The results emerged from the analysis of a new system which appeared due to degeneracy.

    Our study is performed in a general framework and produces generic results which can be applied in particular systems exhibiting this bifurcation. The double-Hopf bifurcation (degenerate or not) can be often met in practical models based on differential systems and having a large number of equations (at least four) [4,9,10,11,12]. Interesting results on discrete time systems undergoing double-Hopf bifurcations are presented in [11]. Thus, understanding better this bifurcation and obtaining new generic properties of it is desirable and important for particular models describing real-world phenomena. This is the purpose of the present work.

    The paper is organized as follows. In Section 2 the classical normal form obtained in [1] using six generic conditions (HH.0)–(HH.5) is presented and a new normal form up to order 3 in polar coordinates is obtained as two of the generic conditions, namely (HH.1) and (HH.3) fail. In Section 3, the dynamics and bifurcation of this normal form, around the origin, is analyzed as the condition (HH.3) is not satisfied, but (HH.1) is valid, while in Section 4 a similar study is done when the condition (HH.1) is not fulfilled, but (HH.3) is satisfied. In Section 5 some conclusions and the relationship between the dynamics of the 2D amplitude system and the dynamics of the 4D truncated normal form are given.

    Consider a differential system of the form

    $ \begin{equation} \dot{x} = f\left( x,\alpha \right) ,\ \ \ x\in \mathbb{R}^{4},\ \ \ \alpha = \left( \alpha _{1},\alpha _{2}\right) \in \mathbb{R}^{2}, \end{equation} $ (2.1)

    with $ f $ smooth, $ \dot{x} = \frac{dx}{dt} $, and assume that $ x = 0 $ is an equilibrium point of the system for all $ \alpha $ with $ \left\vert \alpha \right\vert = \sqrt{\alpha _{1}^{2}+\alpha _{1}^{2}} $ small enough, that is, $ f\left(0, \alpha \right) \equiv 0; $ $ x = 0 $ stands for $ x = (0, 0, 0, 0) $ and $ \alpha = 0 $ for $ \alpha = (0, 0). $ The system (2.1) can be written as

    $ \begin{equation} \dot{x} = A\left( \alpha \right) x+F\left( x,\alpha \right) \end{equation} $ (2.2)

    where $ F\left(x, \alpha \right) = O\left(\left\vert x\right\vert ^{2}\right) $ is a smooth function denoting Taylor rest with terms of order at least $ 2. $

    Assume the matrix $ A(\alpha) $ has two pairs of simple complex-conjugate eigenvalues $ \lambda _{1}, $ $ \overline{\lambda }_{1}, $ $ \lambda _{2}, $ $ \overline{\lambda }_{2}, $

    $ \begin{equation*} \lambda _{1}\left( \alpha \right) = \mu _{1}\left( \alpha \right) +i\omega _{1}\left( \alpha \right) ,\ \ \ \lambda _{2}\left( \alpha \right) = \mu _{2}\left( \alpha \right) +i\omega _{2}\left( \alpha \right) \end{equation*} $

    for all sufficiently small $ \left\vert \alpha \right\vert, $ where $ \mu _{1, 2}\left(\alpha \right) $ and $ \omega _{1, 2}\left(\alpha \right) $ are smooth functions of $ \alpha $ such that $ \mu _{1}\left(0\right) = \mu _{2}\left(0\right) = 0, $ respectively, $ \omega _{1}\left(0\right) = \omega _{10} > 0 $ and $ \omega _{2}\left(0\right) = \omega _{20} > 0. $ When these conditions are satisfied, the origin of system (2.1) is called a double-Hopf or Hopf-Hopf singularity, and, under supplementary conditions, a double-Hopf bifurcation occurs at $ \alpha = 0 $.

    In complex coordinates, the system (2.2) can be further reduced to the form

    $ \begin{equation} \left\{ \begin{array}{c} \dot{z}_{1} = \lambda _{1}\left( \alpha \right) z_{1}+g\left( z_{1},\overline{z }_{1},z_{2},\overline{z}_{2},\alpha \right) \\ \dot{z}_{2} = \lambda _{2}\left( \alpha \right) z_{2}+h\left( z_{1},\overline{z }_{1},z_{2},\overline{z}_{2},\alpha \right) \end{array} \right. \end{equation} $ (2.3)

    where

    $ \begin{equation*} g\left( z_{1},\overline{z}_{1},z_{2},\overline{z}_{2},\alpha \right) = \sum\limits_{j+k+l+m\geq 2}g_{jklm}\left( \alpha \right) z_{1}^{j}\overline{z} _{1}^{k}z_{2}^{l}\overline{z}_{2}^{m} \end{equation*} $

    and

    $ \begin{equation*} h\left( z_{1},\overline{z} _{1},z_{2},\overline{z}_{2},\alpha \right) = \sum\limits_{j+k+l+m\geq 2}h_{jklm}\left( \alpha \right) z_{1}^{j}\overline{z}_{1}^{k}z_{2}^{l} \overline{z}_{2}^{m}. \end{equation*} $

    Other two equations are the conjugates of system $ \left(2.3\right) $ and are omitted in what follows.

    A first normal form corresponding to system(2.3) is described in the following lemma, reported in [1].

    Lemma 2.1. Assume that the following conditions take place [1]

    $ \begin{equation*} \mathit{\mbox{(HH.0)}} \qquad k\omega _{10}\neq l\omega _{20},\ \ \ k,l > 0,\ \ \ k+l\leq 5. \end{equation*} $

    Then, there exists a locally defined, smooth and smoothlyparameter-dependent, invertible transformation of the complex variables thatreduces system $ (2.3) $ for all sufficiently small $ \left\vert \alpha\right\vert $ into the following form:

    $ \begin{equation} \left\{ \begin{array}{c} \dot{w}_{1} = \lambda _{1}w_{1}+w_{1}\left\vert w_{1}\right\vert ^{2}\left( a_{1}+ib_{1}\right) +w_{1}\left\vert w_{2}\right\vert ^{2}\left( a_{2}+ib_{2}\right) +w_{1}\left\vert w_{1}\right\vert ^{4}\left( a_{3}+ib_{3}\right) \\ +w_{1}\left\vert w_{1}\right\vert ^{2}\left\vert w_{2}\right\vert ^{2}\left( a_{4}+ib_{4}\right) +w_{1}\left\vert w_{2}\right\vert ^{4}\left( a_{5}+ib_{5}\right) +O\left( \left\Vert \left( w_{1},\overline{w}_{1},w_{2}, \overline{w}_{2}\right) \right\Vert ^{6}\right) \\ \dot{w}_{2} = \lambda _{2}w_{2}+w_{2}\left\vert w_{1}\right\vert ^{2}\left( c_{1}+id_{1}\right) +w_{2}\left\vert w_{2}\right\vert ^{2}\left( c_{2}+id_{2}\right) +w_{2}\left\vert w_{1}\right\vert ^{4}\left( c_{3}+id_{3}\right) \\ +w_{2}\left\vert w_{1}\right\vert ^{2}\left\vert w_{2}\right\vert ^{2}\left( c_{4}+id_{4}\right) +w_{2}\left\vert w_{2}\right\vert ^{4}\left( c_{5}+id_{5}\right) +O\left( \left\Vert \left( w_{1},\overline{w}_{1},w_{2}, \overline{w}_{2}\right) \right\Vert ^{6}\right) \end{array} \right. \end{equation} $ (2.4)

    where $ w_{1, 2} $ are complex-valued functions, $ \left\Vert \left(w_{1}, \overline{w}_{1}, w_{2}, \overline{w}_{2}\right) \right\Vert ^{2} = \left\vert w_{1}\right\vert ^{2}+\left\vert w_{2}\right\vert ^{2} $.

    In system (2.4) we used the following notations: $ a_{1}+ib_{1} = G_{2100}\left(\alpha \right), $ $ a_{2}+ib_{2} = G_{1011}\left(\alpha \right), $ $ a_{3}+ib_{3} = G_{3200}\left(\alpha \right), $ $ a_{4}+ib_{4} = G_{2111}\left(\alpha \right) $ and $ a_{5}+ib_{5} = G_{1022}\left(\alpha \right), $ respectively, $ c_{1}+id_{1} = H_{1110}\left(\alpha \right), $ $ c_{2}+id_{2} = H_{0021}\left(\alpha \right), $ $ c_{3}+id_{3} = H_{2210}\left(\alpha \right), $ $ c_{4}+id_{4} = H_{1121}\left(\alpha \right) $ and $ c_{5}+id_{5} = H_{0032}\left(\alpha \right), $ where $ a_{i} = a_{i}\left(\alpha \right), $ $ b_{i} = b_{i}\left(\alpha \right), $ $ c_{i} = c_{i}\left(\alpha \right) $ and $ d_{i} = d_{i}\left(\alpha \right), $ $ i = \overline{1, 5}, $ are real smooth functions of $ \alpha $ which depend on the coefficients $ g_{ijkl}\left(\alpha \right) $ and $ h_{ijkl}\left(\alpha \right). $ The coefficients of third order terms in system (2.4) are given explicitly in [1].

    The system (2.4) has been transformed further in [1] to a normal form based on six generic conditions.

    In polar coordinates $ w_{1} = r_{1}e^{i\varphi _{1}} $ and $ w_{2} = r_{2}e^{i \varphi _{2}}, $ the system (2.4) becomes

    $ \begin{equation} \left\{ \begin{array}{l} \dot{r}_{1} & = & r_{1}\left( \mu _{1}+a_{1}r_{1}^{2}+a_{2}r_{2}^{2}+a_{3}r_{1}^{4}+a_{4}r_{1}^{2}r_{2}^{2}+a_{5}r_{2}^{4}\right) +\Phi _{1}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \\ \dot{r}_{2} & = & r_{2}\left( \mu _{2}+c_{1}r_{1}^{2}+c_{2}r_{2}^{2}+c_{3}r_{1}^{4}+c_{4}r_{1}^{2}r_{2}^{2}+c_{5}r_{2}^{4}\right) +\Phi _{2}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \\ \dot{\varphi}_{1} & = & \omega _{1}+\Psi _{1}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \\ \dot{\varphi}_{2} & = & \omega _{2}+\Psi _{2}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \end{array} \right. , \end{equation} $ (2.5)

    where the real functions $ \Phi _{k}, $ $ \Psi _{k} $ are smooth functions of their arguments and are $ 2\pi $-periodic in $ \varphi _{j}, $ $ \Phi _{k} = O\left(\left(r_{1}^{2}+r_{2}^{2}\right) ^{3}\right) $, $ \Psi _{k}\left(0, 0, \varphi _{1}, \varphi _{2}\right) = 0 $.

    The main dynamics of the system (2.5) is given by its first two equations

    $ \begin{equation} \left\{ \begin{array}{l} \dot{r}_{1} = r_{1}\left( \mu _{1}+a_{1}r_{1}^{2}+a_{2}r_{2}^{2}+a_{3}r_{1}^{4}+a_{4}r_{1}^{2}r_{2}^{2}+a_{5}r_{2}^{4}\right) +O\left( \left( r_{1}^{2}+r_{2}^{2}\right) ^{3}\right) \\ \dot{r}_{2} = r_{2}\left( \mu _{2}+c_{1}r_{1}^{2}+c_{2}r_{2}^{2}+c_{3}r_{1}^{4}+c_{4}r_{1}^{2}r_{2}^{2}+c_{5}r_{2}^{4}\right) +O\left( \left( r_{1}^{2}+r_{2}^{2}\right) ^{3}\right) \end{array} \right., \end{equation} $ (2.6)

    the other two describing only the rate of rotation of an orbit, and, because $ \omega _{1, 2}\left(0\right) \neq 0, $ they can be approximated by

    $ \begin{equation*} \dot{\varphi}_{1} = \omega _{10},\ \ \dot{\varphi}_{2} = \omega _{20}. \end{equation*} $

    Changing to $ \rho _{i} = r_{i}^{2}, $ $ i = 1, 2, $ and truncating the higher order terms, system (2.6) leads to the planar system

    $ \begin{equation} \left\{ \begin{array}{l} \dot{\rho_{1}} & = & 2\rho _{1}\left( \mu _{1}+a_{1}\rho _{1}+a_{2}\rho _{2}+a_{3}\rho _{1}^{2}+a_{4}\rho _{1}\rho _{2}+a_{5}\rho _{2}^{2}\right) \\ \dot{\rho_{2}} & = & 2\rho _{2}\left( \mu _{2}+c_{1}\rho _{1}+c_{2}\rho _{2}+c_{3}\rho _{1}^{2}+c_{4}\rho _{1}\rho _{2}+c_{5}\rho _{2}^{2}\right) \end{array} \right. , \end{equation} $ (2.7)

    referred as the amplitude equations; $ \dot{\rho}_{i} = \frac{d\rho _{i}}{dt}, $ $ i = 1, 2. $

    In the hypotheses

    $ \begin{align} \rm{(HH.0)}&\quad k\omega _{1}\left( 0\right) \neq l\omega _{2}\left( 0\right) ,\ \ \ k,l > 0,\ \ \ k+l\leq 5; \\ \rm{(HH.1)} &\quad p_{11}\left( 0\right) = Re\left( G_{2100}\left( 0\right) \right) \neq 0; \\ \rm{(HH.2)} &\quad p_{12}\left( 0\right) = Re\left( G_{1011}\left( 0\right) \right) \neq 0; \\ \rm{(HH.3)} &\quad p_{21}\left( 0\right) = Re\left( H_{1110}\left( 0\right) \right) \neq 0; \\ \rm{(HH.4)} &\quad p_{22}\left( 0\right) = Re\left( H_{0021}\left( 0\right) \right) \neq 0; \\ \rm{(HH.5)}&\quad \rm{the map }\alpha \longrightarrow \mu \left( \alpha \right) \rm{ is regular at }\alpha = 0, \end{align} $ (2.8)

    in [1], the system (2.4) has been transformed further into the following normal form

    $ \begin{equation} \left\{ \begin{array}{l} \dot{r}_{1} & = & r_{1}\left( \mu _{1}+p_{11}r_{1}^{2}+p_{12}r_{2}^{2}+s_{1}r_{2}^{4}\right) +\Phi _{1}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \\ \dot{r}_{2} & = & r_{2}\left( \mu _{2}+p_{21}r_{1}^{2}+p_{22}r_{2}^{2}+s_{2}r_{1}^{4}\right) +\Phi _{2}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \\ \dot{\varphi}_{1} & = & \omega _{1}+\Psi _{1}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \\ \dot{\varphi}_{2} & = & \omega _{2}+\Psi _{2}\left( r_{1},r_{2},\varphi _{1},\varphi _{2},\alpha \right) \end{array} \right. , \end{equation} $ (2.9)

    where the real functions $ \Phi _{k}, $ $ \Psi _{k} $ are smooth functions of their arguments and are $ 2\pi $-periodic in $ \varphi _{j}, $ $ \Phi _{k} = O\left(\left(r_{1}^{2}+r_{2}^{2}\right) ^{3}\right) $, $ \Psi _{k}\left(0, 0, \varphi _{1}, \varphi _{2}\right) = 0. $

    Consequently, the truncated amplitude system obtained in [1] reads:

    $ \begin{equation} \left\{ \begin{array}{l} \dot{\rho}_{1} & = & 2\rho _{1}\left( \mu _{1}+p_{11}\rho _{1}+p_{12}\rho _{2}+s_{1}\rho _{2}^{2}\right) \\ \dot{\rho}_{2} & = & 2\rho _{2}\left( \mu _{2}+p_{21}\rho _{1}+p_{22}\rho _{2}+s_{2}\rho _{1}^{2}\right) \end{array} \right. . \end{equation} $ (2.10)

    This system is analyzed in two cases, the "simple case" as $ p_{11}(0)p_{22}(0) > 0 $ and the "difficult case" as $ p_{11}(0)p_{22}(0) < 0 $.

    The study of the generic Hopf-Hopf bifurcation is also done in [2], starting with the amplitude system (2.6), truncated up to third order terms, in the nondegeneracy hypotheses

    $ \begin{equation} a_{1}(0)\neq 0,a_{2}(0)\neq 0,c_{1}(0)\neq 0,c_{2}(0)\neq 0, \end{equation} $ (2.11)

    and

    $ \begin{equation} a_{1}(0)c_{2}(0)-a_{2}(0)c_{1}(0)\neq 0. \end{equation} $ (2.12)

    In this paper we study the degenerate Hopf-Hopf bifurcation when conditions (HH.1) and (HH.3) are not satisfied. In order to do this, we derive a new normal form for the amplitude system.

    Theorem 2.2. Assume that the following three generic conditions are satisfied:

    $\textrm{(HH.2)}$ $ a_{2}\left(0\right) = Re\left[ G_{1011}\left(0\right) \right] \neq0, $

    $\textrm{(HH.4)}$ $ c_{2}\left(0\right) = Re\left[ H_{0021}\left(0\right) \right] \neq0, $

    $\textrm{(HH.5)}$ the map $ \alpha \mapsto \left(\mu _{1}\left(\alpha \right), \mu_{2}\left(\alpha \right) \right) $ is regular at $ \alpha = 0. $

    Then, system $ (2.6) $ is locally topologically equivalent around theorigin $ O $, for all sufficiently small $ \left\vert \alpha \right\vert $, tothe following system

    $ \begin{equation} \left\{ \begin{array}{l} \frac{d \rho_{1}}{d\tau} & = & 2\rho _{1}\left[ \mu _{1}+p_{11}\rho _{1}+p_{12}\rho _{2}+p_{13}\rho _{1}^{2}+p_{14}\rho _{1}\rho _{2}\right] + \sqrt{\rho _{1}}O\left( \left( \rho _{1}+\rho _{2}\right) ^{3}\right) \\ \frac{d \rho_{2}}{d\tau} & = & 2\rho _{2}\left[ \mu _{2}+p_{21}\rho _{1}+p_{22}\rho _{2}+p_{23}\rho _{1}^{2}+p_{25}\rho _{2}^{2}\right] +\sqrt{\rho _{2}} O\left( \left( \rho _{1}+\rho _{2}\right) ^{3}\right) \end{array} \right. , \end{equation} $ (2.13)

    where

    $ \begin{equation*} p_{11} = a_{1}+\mu _{1}n_{1},p_{12} = a_{2}+\mu _{1}n_{2},p_{13} = a_{3}+a_{1}n_{1},p_{14} = a_{4}+a_{1}n_{2}+a_{2}n_{1} \end{equation*} $

    and

    $ \begin{equation*} p_{21} = c_{1}+\mu _{2}n_{1},p_{22} = c_{2}+\mu _{2}n_{2},p_{23} = c_{3}+c_{1}n_{1},p_{25} = c_{5}+c_{2}n_{2}, \end{equation*} $

    with $ n_{2}\left(\mu \right) = -\frac{a_{5}\left(\mu \right) }{a_{2}\left(\mu\right) } $ and $ n_{1}\left(\mu \right) = -\frac{c_{4}\left(\mu \right)+c_{1}\left(\mu \right) n_{2}\left(\mu \right) }{c_{2}\left(\mu \right) }, $ where $ p_{ij} = p_{ij}\left(\mu \right) $ for $ i = 1, 2 $ and $ j = \overline{1, 5}, $ are well-defined and smooth functions for all $ \left\vert \mu\right\vert $ small enough.

    Proof. We rescale the time by

    $ \begin{equation*} dt = \left( 1+n_{1}\rho _{1}+n_{2}\rho _{2}\right) d\tau \end{equation*} $

    where $ n_{i} = n_{i}\left(\alpha \right), $ $ i = \overline{1, 2}, $ are smooth functions which will be determined later. Then, system (2.6) is orbitally equivalent to the system

    $ \begin{equation*} \left\{ \begin{array}{l} \frac{d\rho _{1}}{d\tau } & = & 2\rho _{1}\left[ \mu _{1}+p_{11} \rho _{1}+p_{12}\rho _{2}+p_{13}\rho _{1}^{2}+p_{14}\rho _{1}\rho _{2}+p_{15} \rho _{2}^{2}\right] +\sqrt{\rho _{1}}O\left( \left\vert \rho \right\vert ^{3}\right) \\ \frac{d\rho _{2}}{d\tau } & = & 2\rho _{2}\left[ \mu _{2}+p_{21} \rho _{1}+p_{22}\rho _{2}+p_{23} \rho _{1}^{2}+p_{24}\rho _{1}\rho _{2}+p_{25}\rho _{2}^{2}\right] +\sqrt{\rho _{2}}O\left( \left\vert \rho \right\vert ^{3}\right) \end{array} \right. , \end{equation*} $

    where $ p_{11} = a_{1}+\mu _{1}n_{1}, $ $ p_{12} = a_{2}+\mu _{1}n_{2}, $ $ p_{13} = a_{3}+a_{1}n_{1} $, $ p_{14} = a_{4}+a_{1}n_{2}+a_{2}n_{1}, $ and $ p_{15} = a_{5}+a_{2}n_{2} $, respectively $ p_{21} = c_{1}+\mu _{2}n_{1}, $ $ p_{22} = c_{2}+\mu _{2}n_{2}, $ $ p_{23} = c_{3}+c_{1}n_{1}, $ $ p_{24} = c_{4}+c_{1}n_{2}+c_{2}n_{1} $ and $ p_{25} = c_{5}+c_{2}n_{2}. $

    Since $ a_{2}\left(0\right) \neq 0 $ and $ c_{2}\left(0\right) \neq 0, $ one can nullify the coefficients $ p_{15} $ and $ p_{24} $ by taking $ n_{2}\left(\alpha \right) = -\frac{a_{5}\left(\alpha \right) }{a_{2}\left(\alpha \right) } $ and $ n_{1}\left(\alpha \right) = -\frac{c_{4}\left(\alpha \right) +c_{1}\left(\alpha \right) n_{2}\left(\alpha \right) }{c_{2}\left(\alpha \right) }. $ Notice that $ n_{1}\left(\alpha \right) $ and $ n_{2}\left(\alpha \right) $ are well-defined for all $ \left\vert \alpha \right\vert $ small enough, and so are $ p_{ij}\left(\alpha \right). $

    Assume the map $ \alpha \mapsto \left(\mu _{1}\left(\alpha \right), \mu _{2}\left(\alpha \right) \right) $ is regular at $ \alpha = 0, $ that is,

    $ \begin{equation*} \left. \frac{\partial \mu _{1}}{\partial \alpha _{1}}\frac{\partial \mu _{2} }{\partial \alpha _{2}}-\frac{\partial \mu _{1}}{\partial \alpha _{2}}\frac{ \partial \mu _{2}}{\partial \alpha _{1}}\right\vert _{\alpha = 0}\neq 0. \end{equation*} $

    From the Inverse Function Theorem this condition ensures that for any $ \mu _{1}, \mu _{2}\in \mathbb{R} $ with $ \left\vert \mu \right\vert = \sqrt{\mu _{1}^{2}+\mu _{2}^{2}} $ small enough, there exist $ \alpha _{1} = \alpha _{1}\left(\mu \right) $ and $ \alpha _{2} = \alpha _{2}\left(\mu \right) $ which are obtained locally from the system

    $ \begin{equation*} \mu _{1,2} = \frac{\partial \mu _{1,2}}{\partial \alpha _{1}}\left( 0,0\right) \alpha _{1}+\frac{\partial \mu _{1,2}}{\partial \alpha _{2}}\left( 0,0\right) \alpha _{2}. \end{equation*} $

    This implies that, one can consider further $ \mu = \left(\mu _{1}, \mu _{2}\right) $ as the parameter of the system. Thus, $ \omega _{1, 2} = \omega _{1, 2}\left(\mu \right), $ $ a_{i} = a_{i}\left(\mu \right), $ $ b_{i} = b_{i}\left(\mu \right), $ $ c_{i} = c_{i}\left(\mu \right), $ $ d_{i} = d_{i}\left(\mu \right), $ $ p_{ij} = p_{ij}\left(\mu \right), $ $ i = 1, 2, $ $ j = \overline{1, 5}. $ Since $ \mu _{1, 2}\left(0\right) = 0, $ the properties $ \omega _{1}\left(0\right) = \omega _{10} > 0 $ and $ \omega _{2}\left(0\right) = \omega _{20} > 0 $ are preserved. The theorem is proved.

    Remark that the coefficients $ p_{ij} $ in system (2.13) are not the same as the ones in system (2.10), but the values $ p_{11}\left(0\right), $ $ p_{12}\left(0\right), $ $ p_{21}\left(0\right), $ $ p_{22}\left(0\right) $ are the same as in system (2.8).

    The truncated 4D system of system (2.13) reads

    $ \begin{equation} \left\{ \begin{array}{l} \frac{d\rho_{1}}{d\tau} = 2\rho _{1}\left[ \mu _{1}+p_{11}\rho _{1}+p_{12}\rho _{2}+p_{13}\rho _{1}^{2}+p_{14}\rho _{1}\rho _{2}\right] , \\ \frac{d\rho_{2}}{d\tau} = 2\rho _{2}\left[ \mu _{2}+p_{21}\rho _{1}+p_{22}\rho _{2}+p_{23}\rho _{1}^{2}+p_{25}\rho _{2}^{2}\right] , \\ \frac{d{\varphi}_{1}}{d\tau} = \omega _{10}, \\ \frac{d{\varphi}_{2}}{d\tau} = \omega _{20}, \end{array} \right. \end{equation} $ (2.14)

    where $ p_{12}(0)p_{22}(0)\neq 0, $ because $ p_{12}(0) = a_{2}(0)\neq 0, \; \ $from (HH.2) and $ p_{22}(0) = c_{2}\left(0\right) \neq 0, $ from (HH.4).

    We aim to tackle in this work the normal form system (2.13), truncated up to third order terms, when $ a_{2}\left(0\right) c_{2}\left(0\right) > 0. $

    Assume $ p_{12}\left(0\right) = a_{2}\left(0\right) < 0 $ and $ p_{22}\left(0\right) = c_{2}\left(0\right) < 0. $ Make the changes

    $ \begin{equation} \xi _{1} = -p_{12}\left( \mu \right) \rho _{1},\xi _{2} = -p_{22}\left( \mu \right) \rho _{2},t = 2\tau . \end{equation} $ (2.15)

    The transformation $ \left(\rho _{1}, \rho _{2}\right) \longmapsto \left(\xi _{1}, \xi _{2}\right) $ is well defined for all $ \left\vert \mu \right\vert $ small enough, because $ p_{12}\left(0\right) \neq 0 $ and $ p_{22}\left(0\right) \neq 0 $ are satisfied from (HH.2) and (HH.4), and it is nonsingular as $ a_{2}\left(0\right) c_{2}\left(0\right) \neq 0. $ Since $ \frac{d\xi _{1}}{dt} = -p_{12}\left(\mu \right) \frac{d\rho _{1}}{ d\tau }\frac{d\tau }{dt} $ and $ \frac{d\xi _{2}}{dt} = -p_{22}\left(\mu \right) \frac{d\rho _{2}}{d\tau }\frac{d\tau }{dt}, $ the form system (2.13) in its lowest terms in $ \left(\xi _{1}, \xi _{2}\right) $, known as the truncated form, becomes $ \ $

    $ \begin{equation} \left\{ \begin{array}{l} \dot{\xi}_{1} & = & \xi _{1}\left[ \mu _{1}-\theta \left( \mu \right) \xi _{1}-\gamma \left( \mu \right) \xi _{2}+M\left( \mu \right) \xi _{1}\xi _{2}+N\left( \mu \right) \xi _{1}^{2}\right] \\ \dot{\xi}_{2} & = & \xi _{2}\left[ \mu _{2}-\delta \left( \mu \right) \xi _{1}-\xi _{2}+S\left( \mu \right) \xi _{1}^{2}+P\left( \mu \right) \xi _{2}^{2}\right] \end{array} \right. , \end{equation} $ (2.16)

    where $ \theta \left(\mu \right) = \frac{p_{11}}{p_{12}}\left(\mu \right), $ $ \gamma \left(\mu \right) = \frac{p_{12}}{p_{22}}\left(\mu \right), $ $ M\left(\mu \right) = \frac{p_{14}}{p_{12}p_{22}}\left(\mu \right), $ $ N\left(\mu \right) = \frac{p_{13}}{p_{12}^{2}}\left(\mu \right), $ $ \delta \left(\mu \right) = \frac{p_{21}}{p_{12}}\left(\mu \right), $ $ S\left(\mu \right) = \frac{p_{23}}{p_{12}^{2}}\left(\mu \right) $ and $ P\left(\mu \right) = \frac{p_{25}}{p_{22}^{2}}\left(\mu \right) $. However, in what follows, some of these expressions are needed only at $ \mu = 0.\ $

    The dot over quantities stands now for the derivatives with respect to the new time. Denote also by $ f $ the smooth vector field associated with system (2.16).

    Remark 1. Due to the transformation system (2.15), the stability of an equilibrium point in the 2D system (2.16) is preserved in the 4D system (2.14); see also [1].

    Remark 2. 1) Notice that $ \theta (0) = \frac{a_{1}\left(0\right) }{ a_{2}\left(0\right) }, $ $ \gamma (0) = \frac{a_{2}\left(0\right) }{ c_{2}\left(0\right) } $ and $ \delta (0) = \frac{c_{1}\left(0\right) }{ a_{2}\left(0\right) } $ are well-defined from (HH.2), (HH.4). Moreover, $ \gamma (0) > 0 $ while $ \theta (0)\delta (0) $ can be $ 0. $

    2) When $ p_{12}\left(0\right) > 0 $ and $ p_{22}\left(0\right) > 0, $ the change $ \xi _{1} = p_{12}\left(\mu \right) \rho _{1}, $ $ \xi _{2} = p_{22}\left(\mu \right) \rho _{2}, $ $ t = -2\tau $ lead to

    $ \begin{equation} \left\{ \begin{array}{l} \dot{\xi}_{1} & = & \xi _{1}\left[ -\mu _{1}-\theta \left( \mu \right) \xi _{1}-\gamma \left( \mu \right) \xi _{2}-M\left( \mu \right) \xi _{1}\xi _{2}-N\left( \mu \right) \xi _{1}^{2}\right] \\ \dot{\xi}_{2} & = & \xi _{2}\left[- \mu _{2}-\delta \left( \mu \right) \xi _{1}-\xi _{2}-S\left( \mu \right) \xi _{1}^{2}-P\left( \mu \right) \xi _{2}^{2}\right] \end{array} \right. , \end{equation} $ (2.17)

    that can be reduced to system (2.16).

    As $ \rho _{i} = r_{i}^{2}\geq 0, $ $ i = 1, 2, $ taking into account system (2.15) and the fact that $ p_{12}\left(0\right) < 0, $ $ p_{22}\left(0\right) < 0, $ it follows that the system (2.16) must be studied only on the set (the first quadrant)

    $ \begin{equation*} D = \left\{ \left( \xi _{1},\xi _{2}\right) \in \mathbb{R}^{2},\xi _{1}\geq 0,\xi _{2}\geq 0\right\} . \end{equation*} $

    Thus, only the equilibria of system (2.16) situated in $ D $ will be analyzed.

    Remark 3. The lines $ \xi _{1} = 0 $ and $ \xi _{2} = 0 $ are invariant curves for system (2.16), thus the set $ D $ is invariant with respect to the dynamical system associated to system (2.16).

    In the following we analyze the truncated amplitude system (2.16) in two cases when the Hopf-Hopf bifurcation degenerates, namely (ⅰ) $ \delta \left(0\right) = 0, $ $ \theta (0)\neq 0; $ (ⅱ) $ \theta \left(0\right) = 0, $ $ \delta (0)\neq 0. $

    Remind that system (2.16) which is analyzed in the next two sections was obtained in the hypotheses (HH.0), (HH.2), (HH.4), (HH.5), and, as $ p_{12}(0) < 0 $, $ p_{22}(0) < 0 $, we have $ \gamma(0) > 0 $.

    In addition, our study is done for $ \mu $ in a neighborhood of the origin $ V_{\varepsilon } $ given by $ \left\vert \mu \right\vert < \varepsilon $, for $ \varepsilon > 0 $ sufficiently small.

    As all the coefficients are smooth functions depending on the parameter $ \mu $, for $ \mu \in V_{\varepsilon } $ we can write

    $ \theta (\mu ) = \theta(0)+ \frac{\partial \theta }{\partial \mu _{1}}\left( 0\right)\mu _{1} +\frac{\partial \theta }{\partial \mu _{2}}\left( 0\right)\mu _{2}+O(\left\vert \mu \right\vert ^{2}), $

    and similar for the other coefficients. In all the expressions below only the significant lower order terms in $ \mu_1, \mu_2 $ will be considered, as necessary. Also, in order to save symbols, if, for instance we have $ N(0)\ne 0 $ we shall denote $ N = N(0), $ thus, close to $ \mu = 0 $, we have $ N(\mu) = N+O(|\mu |), $ and so on for the other coefficients.

    Assume that $ \delta (0) = 0, $ $ \theta (0)\neq 0. $ This means that condition (HH.3) is not satisfied, but (HH.1) is valid.

    Then $ \delta (\mu) = \delta _{1}\mu _{1}+\delta _{2}\mu _{2}+O(\left\vert \mu \right\vert ^{2}), $ with $ \delta _{1} = \frac{ \partial \delta }{\partial \mu _{1}}\left(0\right), $ $ \delta _{2} = \frac{ \partial \delta }{\partial \mu _{2}}\left(0\right). $ One of the equilibria of system (2.16) is $ E_{0} = \left(0, 0\right). $ Two more equilibria

    $ \begin{equation*} E_{1} = \left( \frac{1}{\theta }\mu _{1}+O\left( \mu _{1}^{2}\right) ,0\right) \;{\rm{ and }}\;E_{2} = \left( 0,\mu _{2}+O\left( \mu _{2}^{2}\right) \right) \end{equation*} $

    where $ \theta = \theta(0) $, bifurcate from $ E_{0} $ as soon as $ \mu _{1}\neq 0, $ respectively, $ \mu _{2}\neq 0 $, i.e. at the bifurcation lines

    $ \begin{equation*} Y_{-} = \{(\mu _{1},\mu _{2}),\mu _{1} = 0,\mu _{2} < 0\},\quad Y_{+} = \{(\mu _{1},\mu _{2}),\mu _{1} = 0,\mu _{2} > 0\}, \end{equation*} $

    and

    $ \begin{equation*} X_{-} = \{(\mu _{1},\mu _{2}),\mu _{2} = 0,\mu _{1} < 0\},\quad X_{+} = \{(\mu _{1},\mu _{2}),\mu _{2} = 0,\mu _{1} > 0\}, \end{equation*} $

    respectively. Their stability is described by Lemma 3.1.

    Lemma 3.1. (1) The eigenvalues of $ E_{0} $ are $ \mu _{1} $ and $ \mu _{2}. $

    (2) The eigenvalues of $ E_{1} $ are $ -\mu _{1}+O\left(\mu _{1}^{2}\right) $and $ \left(\mu _{2}+\frac{S-\delta _{1}\theta }{\theta ^{2}}\mu_{1}^{2}\right) \left(1+O\left(\mu _{1}\right) \right) $, thus, whenever $ E_{1} $ lies in $ D, $ $ E_{1} $ is either (i) a saddle as $ \mu _{1}\left(\mu_{2}+\frac{S-\delta _{1}\theta }{\theta ^{2}}\mu _{1}^{2}\right) > 0, $ (ii)an unstable node as $ \mu _{2}+\frac{S-\delta _{1}\theta }{\theta ^{2}}\mu_{1}^{2} > 0, \mu _{1} < 0, \ $(iii) a stable node as $ \mu _{2}+\frac{S-\delta_{1}\theta }{\theta ^{2}}\mu _{1}^{2} < 0, \mu _{1} > 0. $

    (3) The eigenvalues of $ E_{2} $ are $ \mu _{1}-\gamma \mu _{2}+O(\mu _{2}^{2}) $and $ -\mu _{2}+O(\mu _{2}^{2}). $ Therefore, for $ \left\vert \mu \right\vert $sufficiently small, $ \ $ $ E_{2} $ is (i) a saddle as $ \mu _{1}-\gamma \mu_{2} > 0, $ (ii) a stable node as $ \mu _{1}-\gamma \mu _{2} < 0, $ whenever itexists in $ D. $

    On the other hand, system (2.16) has the nontrivial equilibrium $ E_{3} = \left(\xi _{1}^{\ast }, \xi _{2}^{\ast }\right), $ with

    $ \begin{equation*} \left\{ \begin{array}{c} \xi _{1}^{\ast } = \frac{1}{\theta }\mu _{1}\left( 1+O(\left\vert \mu \right\vert \right) -\frac{\gamma }{\theta }\mu _{2}\left( 1+O(\left\vert \mu \right\vert \right) \rm{ }\ \ \ \ \\ \xi _{2}^{\ast } = \mu _{2}\left( 1+O(\left\vert \mu \right\vert \right) + \frac{S-\delta _{1}\theta }{\theta ^{2}}\mu _{1}^{2}\left( 1+O\left( \left\vert \mu \right\vert \right) \right) \end{array} \right. . \end{equation*} $

    The associated eigenvalues satisfy

    $ \begin{eqnarray} \lambda _{1}\lambda _{2} & = &\xi _{1}^{\ast }\xi _{2}^{\ast }\left[ \theta +O\left( \left\vert \mu \right\vert \right) \right] , \\ \lambda _{1}+\lambda _{2} & = &-\mu _{1}+(\gamma -1)\mu _{2}+O\left( \left\vert \mu \right\vert ^{2}\right) . \end{eqnarray} $ (3.1)

    When $ \theta > 0, $ $ E_{3} $ lies in $ D $ for parameters in the region

    $ \begin{equation*} R_{1} = \left\{ \left( \mu _{1},\mu _{2}\right) \in V_{\varepsilon } ,\mu _{1}-\gamma \mu _{2} > 0,\mu _{2}+\frac{S-\delta _{1}\theta }{\theta ^{2}}\mu _{1}^{2} > 0\right\}. \end{equation*} $

    It is easy to see that for $ \mu \in R_{1}, $ with $ \left\vert \mu \right\vert $ sufficiently small, we have $ \lambda _{1}\lambda _{2} > 0 $ and the curve $ \lambda _{1}+\lambda _{2} = 0 $ does not cross region $ R_{1}, $ thus $ \lambda _{1}+\lambda _{2} < 0, $ and $ E_{3} $ is stable.

    When $ \theta < 0, $ $ E_{3} $ lies in $ D $ for parameters in region

    $ \begin{equation*} R_{2} = \left\{ \left( \mu _{1},\mu _{2}\right) \in V_{\varepsilon } ,\mu _{1}-\gamma \mu _{2} < 0,\mu _{2}+\frac{S-\delta _{1}\theta }{\theta ^{2}}\mu _{1}^{2} > 0\right\} . \end{equation*} $

    By system (3.1), it follows $ \lambda _{1}\lambda _{2} < 0, $ for sufficiently small $ \left\vert \mu \right\vert. $ Thus, $ E_{3} $ is a saddle. The following result is proved.

    Lemma 3.2. If $ \theta < 0, $ then $ E_{3} $ is a saddle, while if $ \theta > 0, $ the equilibrium $ E_{3} $ is a hyperbolic attractor, for allsufficiently small $ \left\vert \mu \right\vert, $ for which $ E_{3} $ lies in $ D. $

    Consequently there can be no Hopf bifurcation at $ E_{3}. $

    As the term $ \delta _{2} $ does not influence the topological type of equilibria, we may restrict our attention only to the $ \left(\theta, \delta _{1}\right) - $ plane.

    Regions $ R_{1} $ and $ R_{2} $ are delimitated by the curves

    $ \begin{eqnarray} T_{1} & = &\left\{\left( \mu _{1},\mu _{2}\right) \in V_{\varepsilon }, \mu _{1} = \gamma \mu _{2}+O\left( \mu _{2}^{2}\right) ,\mu _{2} > 0\right\} , \end{eqnarray} $ (3.2)
    $ \begin{eqnarray} \rm{ }T_{2} & = &\left\{ \left( \mu _{1},\mu _{2}\right) \in V_{\varepsilon },\mu _{2} = -\frac{S-\delta _{1}\theta }{\theta ^{2}}\mu _{1}^{2}+O\left( \mu _{1}^{3}\right) \right\} . \end{eqnarray} $ (3.3)

    We notice that $ E_{3} $ collides with $ E_{2} $ on $ T_{1}, $ and it collides with $ E_{1} $ on $ T_{2}. $

    Proposition 3.3. System $ (2.16) $ experiences the following transcritical bifurcations:

    (i) at the point $ E_{0} $ as the parameter $ \mu _{1} $ varies through thebifurcation value $ \mu _{1} = 0, $ for a fixed $ \mu _{2}\neq 0 $ (when $ E_{0} = E_{1}); $

    (ii) at the point $ E_{0} $ as the parameter $ \mu _{2} $ varies through thebifurcation value $ \mu _{2} = 0, $ for a fixed $ \mu _{1}\neq 0 $ (when $ E_{0} = E_{2}); $

    (iii) at the point $ E_{1} $ as the parameter $ (\mu _{1}, \mu _{2}) $ crossesthe curve $ T_{2} $ (when $ E_{1} = E_{3}); $

    (iv) at the point $ E_{2} $ as the parameter $ (\mu _{1}, \mu _{2}) $ crosses thecurve $ T_{1} $ (when $ E_{2} = E_{3}). $

    Proof. We apply Sotomayor Theorem ([13,14], to prove these statements.

    (ⅰ) The Jacobian matrix $ Df(E_{0}, \mu _{0}) $ at $ \mu _{0} = \left(0, \mu _{2}\right), $ $ \mu _{2}\neq 0, $ has a zero eigenvalue with the right eigenvector $ v = \left(1, 0\right) ^{T} $ and the left eigenvector $ w = \left(1, 0\right) ^{T}. $ It follows

    $ \begin{equation*} w^{T}f_{\mu _{1}}\left( E_{0},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{1}}(E_{0},\mu _{0}) = 1\neq 0,\quad w^{T}[D^{2}f(E_{0},\mu _{0})(v,v)] = -2\theta \neq 0, \end{equation*} $

    thus the transcritical bifurcation conditions are satisfied.

    (ⅱ) The Jacobian matrix $ Df(E_{0}, \mu _{0}) $ at $ \mu _{0} = \left(\mu _{1}, 0\right), $ $ \mu _{1}\neq 0, $ has a zero eigenvalue with the right eigenvector $ v = \left(0, 1\right) ^{T} $ and the left eigenvector $ w = \left(0, 1\right) ^{T}. $ It follows

    $ \begin{equation*} w^{T}f_{\mu _{2}}\left( E_{0},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{2}}(E_{0},\mu _{0}) = 1\neq 0,\quad w^{T}[D^{2}f(E_{0},\mu _{0})(v,v)] = -2\neq 0, \end{equation*} $

    ensuring the existence of a transcritical bifurcation.

    (ⅲ) Consider $ \mu _{0}\in T_{2}, $ $ \mu _{1}\neq 0, $ and $ \mu _{2} $ as a bifurcation parameter. We find that $ v = (-\gamma, \theta)^{T} $ and $ w = (0, 1)^{T} $ are right and left eigenvectors of the Jacobian matrix $ Df(E_{1}, \mu _{0}), $ respectively, corresponding to the zero eigenvalue, and

    $ \begin{align*} &w^{T}f_{\mu _{2}}\left( E_{1},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{2}}(E_{1},\mu _{0}) = \theta \neq 0,\\ & w^{T}[D^{2}f(E_{1},\mu _{0})(v,v)] = 2\theta (\delta \gamma -\theta )-4S\gamma \mu _{1}\neq 0, \end{align*} $

    consequently, for sufficiently small $ \left\vert \mu \right\vert, \ $the conditions are satisfied.

    (ⅳ) Finally, consider $ \mu _{0}\in T_{1}, $ $ \mu _{2}\neq 0, $ and $ \mu _{1} $ as a bifurcation parameter, thus $ \mu _{0} = (\gamma \mu _{2}, \mu _{2}) $. We find the eigenvectors $ v = (1, -\delta)^{T} $ and $ w = (1, 0)^{T}, $ and

    $ \begin{align*} &w^{T}f_{\mu _{1}}\left( E_{2},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{1}}(E_{2},\mu _{0}) = 2(\delta \gamma -\theta )+2M\mu _{2}\neq 0,\\ & w^{T}[D^{2}f(E_{1},\mu _{0})(v,v)] = 1\neq 0, \end{align*} $

    for sufficiently small $ \left\vert \mu \right\vert. $

    For a fixed $ \gamma > 0, $ and $ S > 0, $ the curves $ S-\delta _{1}\theta = 0, $ $ \theta = 0, $ determine four regions in the $ \left(\theta, \delta _{1}\right) - $ plane, illustrated in Figure 1, corresponding to the following cases:

    Figure 1.  Four regions in the $ (\theta, \delta_1) $ plane, $ \gamma > 0 $, and $ S > 0 $.

    Ⅰ$ _{\delta } $: $ \theta > 0, $ $ S-\delta _{1}\theta < 0; $

    Ⅱ$ _{\delta } $: $ \theta > 0, \ $ $ S-\delta _{1}\theta > 0; $

    Ⅲ$ _{\delta } $: $ \theta < 0, \ $ $ S-\delta _{1}\theta > 0; $

    Ⅳ$ _{\delta } $: $ \theta < 0, \ $ $ S-\delta _{1}\theta < 0. $

    For each region Ⅰ$ _{\delta } $-Ⅳ$ _{\delta }, $ in the parametric portraits in the $ \left(\mu _{1}, \mu _{2}\right) $- plane, the parameter strata are determined by the origin and the bifurcation curves $ X_{-}, $ $ X_{+}, $ $ Y_{-}, $ $ Y_{+}, $ $ T_{1}, $ and $ T_{2}. $ Consequently, the following result is obtained.

    Theorem 3.4. For all $ \gamma > 0, S > 0, $ in the $ \left(\theta, \delta_{1}\right) - $ plane, the bifurcation curves consist of

    $ \begin{equation*} O\cup T_{1}\cup T_{2}\cup X_{-}\cup X_{+}\cup Y_{-}\cup Y_{+}. \end{equation*} $

    The four parameter portraits for $ (\theta, $ $ \delta _{1}) $ in regions Ⅰ$ _{\delta } $, Ⅱ$ _{\delta } $, Ⅲ$ _{\delta } $, Ⅳ$ _{\delta } $ are shown in Figure 2. The 14 generic phase portraits are given in Figure 3.

    Figure 2.  Parametric portraits in the case $ \delta = 0 $.
    Figure 3.  Generic phase portraits in the case $ \delta = 0 $. Legend: a black disc for an attractor, a black square for a repeller, a diamond for a saddle equilibrium.

    In Figure 3 we used the following markers to emphasize the topological type of the equilibria: a black disc for an attractor, a black square for a repeller and a diamond for a saddle point. All of these phase portraits are also found in the nondegenerate double-Hopf bifurcation case [1,2]. As $ S\leq 0, $ a similar study can be done.

    Remark 4. As proved in Proposition 3.3, each of the curves $ X_{-}, $ $ X_{+}, $ $ Y_{-}, $ $ Y_{+}, $ $ T_{1}, $ $ T_{2} $ consists of values of transcritical bifurcation (for which two of the four equilibria collide and change topological type), separating two parameter strata, one where both equilibria involved in the transcritical bifurcation lie on $ D $ and one where only one of them lies on $ D. $ Although for parameters on a bifurcation curve the corresponding equilibrium is a saddle-node, the generic phase portrait is equivalent to the one of the stratum where only one of the two equilibria is in $ D. $

    Assume $ \theta (0) = 0, $ $ \delta (0)\neq 0. $ In this case the condition (HH.1) is not satisfied, and (HH.3) is valid. Then $ \theta (\mu) = \theta _{1}\mu _{1}+\theta _{2}\mu _{2}+O(\left\vert \mu \right\vert ^{2}). $

    In this case, two trivial equilibria of system (2.16) are $ E_{0} = \left(0, 0\right), $ with the eigenvalues $ \mu _{1} $ and $ \mu _{2}, $ and $ E_{2} = \left(0, \mu _{2}+O\left(\mu _{2}^{2}\right) \right) $ with eigenvalues $ \mu _{1}-\gamma \mu _{2}+O(\left\vert \mu\right\vert ^2) $ and $ -\mu _{2} +P\mu _{2}^{2}(1+O(\left\vert \mu\right\vert)) $. Remark that these equilibria keep the same form as in the case treated in the previous section, and their stability remains as described by Lemma 3.1.

    As $ \xi _{2} = 0, $ the equation giving the equilibrium points reads:

    $ \begin{equation} \mu _{1}-\theta (\mu) \xi _{1}+N (\mu)\xi _{1}^{2} = 0. \end{equation} $ (4.1)

    Assume that $ N(0)\neq 0, $ thus $ N\left(\mu \right) \neq 0 $ for sufficiently small $ \left\vert \mu \right\vert. $ If $ \Delta (\mu) = \theta ^{2}(\mu)-4N(\mu)\mu _{1}\geq 0, $ Eq (4.1) has two real solutions

    $ \begin{equation*} \xi _{11}\left( \mu \right) = \frac{\theta \left( \mu \right) -\sqrt{\Delta \left( \mu \right) }}{2N\left( \mu \right) },\quad \xi _{12}\left( \mu \right) = \frac{\theta \left( \mu \right) +\sqrt{\Delta \left( \mu \right) }}{ 2N\left( \mu \right) },\quad \xi _{11}\leq \xi _{12}. \end{equation*} $

    Thus, as $ \mu _{1}N > 0, $ two equilibrium points, $ E_{11} = \left(\xi _{11}, 0\right) $ and $ E_{12} = \left(\xi _{12}, 0\right), $ are lying on the positive $ \xi _{1}- $axis, provided that $ \left(\mu _{1}, \mu _{2}\right) \in R_{3}, $ where $ R_{3} $ is a region in the parametric plane given by

    $ \begin{equation} R_{3} = \left\{ \left( \mu _{1},\mu _{2}\right) \in V_{\varepsilon }\left\vert \Delta \left( \mu \right) > 0,\mu _{1}N > 0,\theta N > 0\right. \right\} . \end{equation} $ (4.2)

    Obviously, as $ \Delta \left(\mu \right) = 0, $ we have $ E_{11} = E_{12} = \frac{ \theta }{2N}. $

    As $ \mu _{1}N < 0, $ we have $ \xi _{11} < 0 < \xi _{12} $, thus, only the equilibrium $ E_{12} $ lies in D for parameters in region

    $ \begin{equation} R_{4} = \left\{ \left( \mu _{1},\mu _{2}\right) \in V_{\varepsilon }\left\vert \Delta (\mu ) > 0,\mu _{1}N < 0\right. \right\} . \end{equation} $ (4.3)

    As $ \mu _{1} = 0, $ we have $ E_{11} = E_{0} $ if $ \theta _{2}\mu _{2}N > 0, $ while $ E_{12} = E_{0} $ if $ \theta _{2}\mu _{2}N < 0. $ Using the Implicit Functions Theorem, we find that, for sufficiently small $ \left\vert \mu \right\vert, $ $ \Delta \left(\mu \right) = 0 $ for parameters on the curve, denoted also by $ \Delta, $ given as

    $ \begin{equation*} \Delta = \left\{ \left( \mu _{1},\mu _{2}\right) \in V_{\varepsilon },\mu _{1} = \frac{\theta _{2}^{2}}{4N}\mu _{2}^{2}+O(\mu _{2}^{3}),\theta _{2}\mu _{2}N > 0\right\} . \end{equation*} $

    The eigenvalues of $ E_{11} = \left(\xi _{11}, 0\right) $ satisfy $ \lambda _{1}^{E_{11}} = -\xi _{11}\sqrt{\Delta \left(\mu \right) }\leq 0, $ and

    $ \begin{equation*} \lambda _{2}^{E_{11}} = \mu _{2}-\frac{S}{N}\mu _{1}+\xi _{11}\left( \frac{ \theta S}{N}-\delta \right) . \end{equation*} $

    For $ E_{12} = \left(\xi _{12}, 0\right) $ we have $ \lambda _{1}^{E_{12}} = \xi _{12}\sqrt{\Delta \left(\mu \right) }\geq 0 $ and

    $ \begin{equation*} \lambda _{2}^{E_{12}} = \mu _{2}-\frac{S}{N}\mu _{1}+\xi _{12}\left( \frac{\theta S}{N}-\delta \right) . \end{equation*} $

    An important relation for studying the behavior of $ E_{11} $ and $ E_{12} $ is

    $ \begin{equation} \lambda _{2}^{E_{11}}\lambda _{2}^{E_{12}} = \frac{1}{N}\delta ^{2}\mu _{1}(1+O(\left\vert \mu \right\vert ))+\frac{1}{N}(N-\theta _{2}\delta )\mu _{2}^{2}(1+O(\left\vert \mu \right\vert ). \end{equation} $ (4.4)

    Note that as $ \lambda _{2}^{E_{11}}\lambda _{2}^{E_{12}} = 0, $ we get

    $ \begin{equation*} \lambda _{2}^{E_{11}}+\lambda _{2}^{E_{12}} = \frac{1}{N}(2N-\theta _{2}\delta )\mu _{2}+O(\mu _{2}^{2}). \end{equation*} $

    Lemma 4.1. As the parameter $ \mu $ crosses the curve $ \Delta, $ asaddle-node bifurcation takes place. In addition, if $ N > 0, $ for parameter sufficiently small, close to the curve $ \Delta, $ we have:

    (i) $ E_{11} $ is an attractor and $ E_{12} $ is a saddle as $ \mu _{2}\left(2N-\delta \theta _{2}\right) < 0; $

    (ii) $ E_{11} $ is a saddle and $ E_{12} $ is a repeller as $ \mu _{2}\left(2N-\delta \theta _{2}\right) > 0. $

    Proof. Consider $ \mu _{0}\in \Delta. $ Then $ \xi _{11} = \xi _{12} = \frac{ \theta _{2}}{2N}\mu _{2}+O(\mu _{2}^{2}), $ and the eigenvalues of equilibrium $ E_{11} $ are $ \lambda _{1} = 0 $, $ \lambda _{2} = \frac{2N-\delta \theta _{2}}{2N}\mu _{2}+O(\mu _{2}^{2}). $ The Jacobian matrix $ Df(E_{11}, \mu _{0}), $ has for the zero eigenvalue the right eigenvector $ v = \left(1, 0\right) ^{T} $ and the left eigenvector $ w = \left(2N-\delta \theta _{2}, \gamma \theta _{2}\right) ^{T}. $ It follows

    $ \begin{equation*} w^{T}f_{\mu _{1}}\left( E_{0},\mu _{0}\right) = \frac{\theta _{2}}{2N}\mu _{2}\neq 0,\quad w^{T}[D^{2}f(E_{0},\mu _{0})(v,v)] = \theta _{2}\mu _{2}+O(\mu _{2}^{2})\neq 0, \end{equation*} $

    for sufficiently small $ \left\vert \mu \right\vert. $ Thus, according to Sotomayor Theorem, a saddle-node bifurcation takes place.

    For parameters close to the bifurcation curve we have $ sign(\lambda _{1}^{E_{11}}) = -sign(N), $ $ sign(\lambda _{1}^{E_{12}}) = sign(N), $ and $ sign(\lambda _{2}^{E_{11}}) = sign(\lambda _{2}^{E_{12}}) = sign(\frac{2N-\delta \theta _{2}}{2N}\mu _{2}). $ Consequently, if $ N > 0, $ the equilibrium $ E_{11} $ is an attractor and $ E_{12} $ is a saddle as $ \mu _{2}\left(2N-\delta \theta _{2}\right) < 0, $ while as $ \mu _{2}\left(2N-\delta \theta _{2}\right) > 0, $ the equilibrium $ E_{11} $ is a saddle and $ E_{12} $ is a repeller.

    Denote by

    $ \begin{equation*} \Delta _{+} = \left\{ \left( \mu _{1},\mu _{2}\right) \in \Delta ,\mu _{2}\left( 2N-\delta \theta _{2}\right) > 0\right\} \end{equation*} $

    and

    $ \begin{equation*} \Delta _{-} = \left\{ \left( \mu _{1},\mu _{2}\right) \in \Delta ,\mu _{2}\left( 2N-\delta \theta _{2}\right) < 0\right\} . \end{equation*} $

    As a consequence of Lemma 4.1, for a parameter $ \mu $ in $ \Delta _{+}, $ sufficiently small, the equilibrium $ E_{11}( = E_{12}) $ is a saddle-node, with one attractive direction and three repelling directions, while for a parameter $ \mu $ in $ \Delta _{-}, $ the equilibrium $ E_{11} $ is a saddle-node, with one repelling direction and three attractive ones.

    Consider the curves

    $ \begin{eqnarray*} Y_{-}^{\prime } & = &\left\{ \left( \mu _{1},\mu _{2}\right) \in V_\varepsilon,\mu _{1} = 0,\theta _{2}N\mu _{2} < 0\right\} , \\ Y_{+}^{\prime } & = &\left\{ \left( \mu _{1},\mu _{2}\right) \in V_\varepsilon ,\mu _{1} = 0,\theta _{2}N\mu _{2} > 0\right\} . \end{eqnarray*} $

    Proposition 4.2. The following statements hold:

    (i) for $ \mu \in Y_{+}^{\prime }, $ equilibria $ \ E_{0} $ and $ E_{11} $coincide; when parameters cross this curve, a transcritical bifurcationtakes place, and $ E_{0} $ and $ E_{11} $ interchange the topological type;

    (ii) for $ \mu \in Y_{-}^{\prime }, $ equilibria $ \ E_{0} $ and $ E_{12}\ $coincide and are saddle-nodes. A transcritical bifurcation takes place whenparameters are crossing the curve $ Y_{-}^{\prime }, $ and $ E_{0} $ and $ E_{12} $interchange the topological type.

    (iii) for $ \mu \in \left\{ \left(\mu _{1}, \mu _{2}\right), \mu _{2} = 0, \mu_{1}\neq 0\right\}, $ equilibria $ \ E_{0} $ and $ E_{2}\ $coincide and atranscritical bifurcation takes place when parameters are crossing thiscurve.

    Proof. As in Proposition 3.3, we apply Sotomayor Theorem ([13]) to prove these statements.

    (ⅰ) Consider $ \mu _{0}\in Y_{+}^{\prime }. $ As $ \mu _{1} = 0 $ and $ \theta _{2}\mu _{2}N > 0, $ it follows that $ E_{11} = E_{0} $. The Jacobian matrix $ Df(E_{0}, \mu _{0}) $ at $ \mu _{0} = \left(0, \mu _{2}\right), $ $ \mu _{2}\neq 0, $ has a zero eigenvalue with the right eigenvector $ v = \left(1, 0\right) ^{T} $ and the left eigenvector $ w = \left(1, 0\right) ^{T}. $ It follows

    $ \begin{align*} &w^{T}f_{\mu _{1}}\left( E_{0},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{1}}(E_{0},\mu _{0}) = 1\neq 0,\\ & w^{T}[D^{2}f(E_{0},\mu _{0})(v,v)] = -2\theta _{2}\mu _{2}\neq 0, \end{align*} $

    for sufficiently small $ \left\vert \mu \right\vert, $ thus the transcritical bifurcation conditions are satisfied.

    (ⅱ) Consider $ \mu _{0}\in Y_{-}^{\prime }. $ As $ \mu _{1} = 0 $ and $ \theta _{2}\mu _{2}N < 0, $ it follows that $ \xi _{12} = 0, $ thus $ E_{12} = E_{0} $. We obtain the same values for the quantities involved in the transcritical bifurcation conditions as in case (i).

    (ⅲ) The Jacobian matrix $ Df(E_{0}, \mu _{0}) $ at $ \mu _{0} = \left(\mu _{1}, 0\right), $ $ \mu _{1}\neq 0, $ has a zero eigenvalue with the right eigenvector $ v = \left(0, 1\right) ^{T} $ and the left eigenvector $ w = \left(0, 1\right) ^{T}. $ It follows

    $ \begin{equation*} w^{T}f_{\mu _{2}}\left( E_{0},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{2}}(E_{0},\mu _{0}) = 1\neq 0,\quad w^{T}[D^{2}f(E_{0},\mu _{0})(v,v)] = -2\neq 0, \end{equation*} $

    ensuring the existence of a transcritical bifurcation.

    For $ E_{3} = \left(\xi _{1}^{\ast }, \xi _{2}^{\ast }\right), $ solution of system

    $ \begin{equation*} \left\{ \begin{array}{c} \mu _{1}-(\theta _{1}\mu _{1}+\theta _{2}\mu _{2})\xi _{1}-\gamma (\mu) \xi _{2}+M(\mu)\xi _{1}\xi _{2}+N(\mu)\xi _{1}^{2} = 0, \\ \mu _{2}-\delta(\mu) \xi _{1}-\xi _{2}+S(\mu)\xi _{1}^{2}+P(\mu) \xi _{2}^{2} = 0, \end{array} \right. \end{equation*} $

    we obtain

    $ \begin{eqnarray*} \xi _{1}^{\ast } & = &-\frac{1}{\gamma \delta }(\mu _{1}-\gamma \mu _{2})+O(\left\vert \mu \right\vert ^{2}), \\ \xi _{2}^{\ast } & = &\frac{1}{\gamma }\mu _{1}(1+O(\left\vert \mu \right\vert ))+\frac{N-\theta _{2}\delta }{\delta ^{2}\gamma }\mu _{2}^{2}\left( 1+O\left( \left\vert \mu \right\vert \right) \right) . \end{eqnarray*} $

    Thus, $ E_{3} $ is a nontrivial equilibrium for system (2.16) provided

    $ \begin{equation*} \mu \in R_{5} = \left\{ \left( \mu _{1},\mu _{2}\right) \in V_\varepsilon ,\delta (\mu _{1}-\gamma \mu _{2}) < 0,\delta ^{2}\mu _{1}+\left( N-\theta _{2}\delta \right) \mu _{2}^{2} > 0\right\} . \end{equation*} $

    Denote by

    $ \begin{equation*} T_{3} = \left\{ \left( \mu _{1},\mu _{2}\right) \in V_\varepsilon ,\mu _{1} = \frac{\theta _{2}\delta -N}{\delta ^{2}}\mu _{2}^{2}+O(\mu _{2}^{3})\right\} \ \end{equation*} $

    the parameter curve for which $ \xi _{2}^{\ast } = 0. $ Thus $ R_{5} $ is delimited by $ T_{1} $ and $ T_{3} $.

    The eigenvalues of $ E_{3} $ satisfy

    $ \begin{equation} \lambda _{1}\lambda _{2} = \xi _{1}^{\ast }\xi _{2}^{\ast }\left( -\gamma \delta +O\left( \left\vert \mu \right\vert \right) \right) \;{\rm{ and }}\; \lambda _{1}+\lambda _{2} = \frac{1}{\gamma \delta ^{2}}p\left( \mu \right) , \end{equation} $ (4.5)

    where $ p(\mu) = -\delta ^{2}\mu _{1}\left(1+O\left(\left\vert \mu \right\vert \right) \right) +k_{1}\mu _{2}^{2}\left(1+O\left(\left\vert \mu \right\vert \right) \right) $ and $ k_{1} = \left(\theta _{2}\delta -N\right) +\gamma \left(2N-\theta _{2}\delta \right). $

    Applying Implicit Functions Theorem to the implicit equation $ p\left(\mu _{1}, \mu _{2}\right) = 0, $ the following lemma is obtained.

    Lemma 4.3. If $ (\theta _{2}\delta -N)+\gamma (2N-\theta _{2}\delta)\neq0, $ there exists a neighborhood $ U $ of $ 0 $ in the parameter plane such that $ p\left(\mu \right) = 0 $ in $ U $ iff $ \mu \in U\cap H, $ where

    $ \begin{equation} H = \left\{ \left( \mu _{1},\mu _{2}\right) ,\mu _{1} = \frac{(\theta _{2}\delta -N)+\gamma (2N-\theta _{2}\delta )}{\delta ^{2}}\mu _{2}^{2}+O\left( \mu _{2}^{3}\right) \right\} \end{equation} $ (4.6)

    for $ \left\vert \mu \right\vert $ sufficiently small.

    As a consequence we obtain the following result concerning the topological type of the nontrivial equilibrium $ E_{3}. $

    Lemma 4.4. Assume $ N > 0. $ For sufficiently small $ \left\vert \mu\right\vert, $ the following statements hold:

    (1) If $ \delta > 0, $ then $ E_{3} $ is a saddle.

    (2) If $ \delta < 0, $ $ \theta _{2}\delta -2N\geq 0, $ then $ E_{3} $ is ahyperbolic attractor.

    (3) If $ \delta < 0, $ $ \theta _{2}\delta -2N < 0, $ then the equilibrium $ E_{3} $is a hyperbolic repeller as $ \delta ^{2}\mu _{1} < [\gamma (2N-\theta_{2}\delta)+(\theta _{2}\delta -N)]\mu _{2}^{2}, $ and a hyperbolicattractor for the other parameters in $ R_{5}\backslash H, $

    A similar result can be obtained if $ N < 0. $ We explore next the local bifurcations at the equilibrium $ E_{3}. $

    Proposition 4.5. The following statements hold:

    (i) For $ \mu _{0}\in T_{1}, $ equilibria $ E_{2} $ and $ E_{3} $ coincide. Whenparameter $ \mu $ crosses $ T_{1} $ a transcritical bifurcation takes place at $ E_{3}.$

    (ii) For $ \mu _{0}\in T_{3}, $ $ E_{3} $ coincides either with $ E_{11} $ if $ \frac{2N-\theta _{2}\delta }{2N\delta }\mu _{2} < 0, $ or with $ E_{12} $ if $ \frac{2N-\theta _{2}\delta }{2N\delta }\mu _{2} > 0. $

    (iii) Assume $ 2N-\delta \theta _{2}\neq 0, $ and $ \frac{\partial \delta }{\partial \mu _{1}}\neq 0. $ Then, when parameter $ \mu $ crosses $ T_{3}, $ atranscritical bifurcation takes place at $ E_{3}. $

    Proof. (ⅰ) Consider $ \mu _{0}\in T_{1}, $ $ \mu _{0} = (\gamma \mu _{2}+O(\mu _{2}^{2}), \mu _{2}), $ with $ \mu _{2} > 0. $ Applying Sotomayor theorem, we obtain the right eigenvector $ v = \left(1, -\delta \right) ^{T} $ and the left eigenvector $ w = \left(1, 0\right) ^{T}, \ $and

    $ \begin{align*} &w^{T}f_{\mu _{1}}\left( E_{3},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{1}}(E_{3},\mu _{0}) = 2\gamma \delta +O(\mu _{2})\neq 0,\\ & w^{T}[D^{2}f(E_{0},\mu _{0})(v,v)] = 1\neq 0, \end{align*} $

    for $ \left\vert \mu _{0}\right\vert $ sufficiently small. Thus the transcritical bifurcation conditions are satisfied.

    (ⅱ) Consider $ \mu _{0}\in T_{3}, $ sufficiently small. Then $ \mu _{0} = \left(\frac{\theta _{2}\delta -N}{\delta ^{2}}\mu _{2}^{2}+O(\mu _{2}^{3}), \mu _{2}\right), $ $ \mu _{2}\neq 0. $ We obtain $ \xi _{1}^{\ast } = \frac{1}{\delta }\mu _{2}+O(\mu _{2}^{2})\neq 0, $ $ \xi _{2}^{\ast } = 0, $ and

    $ \begin{equation*} \xi _{1}^{\ast }-\frac{\theta _{1}\mu _{1}+\theta _{2}\mu _{2}}{2N} = \frac{ 2N-\theta _{2}\delta }{2N\delta }\mu _{2}+O(\mu _{2}^{2}), \end{equation*} $

    consequently, $ E_{3} = E_{12} $ if $ \frac{2N-\theta _{2}\delta }{2N\delta }\mu _{2} > 0, $ and $ E_{3} = E_{11} $ if $ \frac{2N-\theta _{2}\delta }{2N\delta }\mu _{2} < 0. $

    (ⅲ) Consider $ \mu _{0}\in T_{3}, $ sufficiently small. In order to prove the existence of a transcritical bifurcation, we choose $ \mu _{1} $ as the bifurcation parameter. Write $ \delta $ as

    $ \begin{equation*} \delta (\mu ) = \delta (0)+\delta _{1}\mu _{1}+\delta _{2}\mu _{2}+O(\left\vert \mu \right\vert ^{2}). \end{equation*} $

    The Jacobian matrix $ Df(E_{3}, \mu _{0}) $ has a zero eigenvalue with the right eigenvector

    $ \begin{equation*} v = \left( \gamma \delta -M\mu _{2}+O(\mu _{2}^{2}),(2N-\theta _{2}\delta )\mu _{2}+O(\mu _{2}^{2})\right) ^{T} \end{equation*} $

    and the left eigenvector $ w = \left(0, 1\right) ^{T}. $ It follows

    $ \begin{equation*} w^{T}f_{\mu _{1}}\left( E_{3},\mu _{0}\right) = 0,\quad w^{T}Df_{\mu _{1}}(E_{3},\mu _{0}) = \frac{\delta _{1}}{\delta }(2N-\delta \theta _{2})\mu _{2}^{2}\left( 1+O\left( \mu _{2}\right) \right) \neq 0, \end{equation*} $
    $ \begin{equation*} w^{T}[D^{2}f(E_{3},\mu _{0})(v,v)] = 2\gamma \delta ^{2}(2N-\delta \theta _{2})\mu _{2}\left( 1+O\left( \mu _{2}\right) \right) \neq 0, \end{equation*} $

    for sufficiently small $ \left\vert \mu \right\vert, $ ensuring the existence of a transcritical bifurcation.

    From Lemma 4.4 it follows that system (2.16) may exhibit a Hopf bifurcation at $ E_{3} $.

    Remark 5. The above analysis shows that the term $ \theta _{1} $ does not influence the topological type of the equilibria, and only the coefficients $ \gamma, \delta, \theta _{2} $ and $ N $ are significant, for sufficiently small $ \left\vert \mu \right\vert. $

    For a fixed $ \gamma > 0, $ and $ N > 0, $ the curves $ \delta = 0, $ $ \theta _{2} = 0, \theta _{2}\delta -N = 0, $ $ \theta _{2}\delta -2N = 0\ $determine eight regions in the $ \left(\theta _{2}, \delta \right) - $ plane (see Figure 4), corresponding to the following cases:

    Figure 4.  Eight regions in the $ (\theta_2, \delta) $ plane, $ \gamma > 0 $, and $ N > 0 $.

    Ⅰ$ _{\theta } $: $ \theta _{2} > 0, \delta > 0, $ $ \theta _{2}\delta -2N > 0; $

    Ⅱ$ _{\theta } $: $ \theta _{2} > 0, \delta > 0, \ \theta _{2}\delta -N > 0, $ $ \theta _{2}\delta -2N < 0; $

    Ⅲ$ _{\theta } $: $ \theta _{2} > 0, \delta > 0, $ $ \theta _{2}\delta -N < 0; $

    Ⅳ$ _{\theta } $: $ \theta _{2} < 0, \delta > 0, $ $ \theta _{2}\delta -N < 0; $

    Ⅴ$ _{\theta } $: $ \theta _{2} < 0, \delta < 0, $ $ \theta _{2}\delta -2N > 0; $

    Ⅵ$ _{\theta } $: $ \theta _{2} < 0, \delta < 0, $ $ \theta _{2}\delta -N > 0, \theta _{2}\delta -2N < 0; $

    Ⅶ$ _{\theta } $: $ \theta _{2} < 0, \delta < 0, $ $ \theta _{2}\delta -N < 0; $

    Ⅷ$ _{\theta } $: $ \theta _{2} > 0, \delta < 0, $ $ \theta _{2}\delta -N < 0. $

    Theorem 4.6. For all $ \gamma > 0, N > 0, $ respectively, $ (\theta _{2}, \delta) $in regions I$ _{\theta } $, II$ _{\theta } $, III$ _{\theta } $, IV$ _{\theta } $, V$ _{\theta } $ in the $ \left(\theta _{2}, \delta \right) - $ plane, thebifurcation curves consist of

    $ \begin{equation*} O\cup T_{1}\cup T_{3}\cup \Delta _{\pm }\cup X_{-}\cup X_{+}\cup Y_{-}^{\prime }\cup Y_{+}^{\prime }. \end{equation*} $

    The generic parameter portraits are shown in Figures 5 and 7, and the corresponding generic phase portraits in Figures 6 and 8.

    Figure 5.  Parametric portraits in the case $ \theta = 0 $, regions corresponding to $ \delta > 0 $.
    Figure 6.  Generic phase portraits in the case $ \theta = 0 $, regions corresponding to $ \delta > 0 $.
    Figure 7.  Parametric portraits in the case $ \theta = 0 $, regions corresponding to $ \delta < 0 $.
    Figure 8.  Generic phase portraits in the case $ \theta = 0 $, regions corresponding to $ \delta < 0 $.

    A Hopf bifurcation may occur when $ \left(\mu _{1}, \mu _{2}\right) $ crosses the curve $ H $ and $ \left(\theta _{2}, \delta \right) $ is situated in regions VI$ _{\theta } $, VII$ _{\theta } $, VIII$ _{\theta }. $ In the case when the first Lyapunov is non-zero, a limit cycle surrounding the equilibrium $ E_{3} $ appear. This cycle is stable if the Hopf bifurcation is supercritical (the first Lyapunov coefficient is negative) or is unstable if the Hopf bifurcation is subcritical (i.e., the first Lyapunov coefficient is positive). As the first Lyapunov coefficient is zero, the equilibrium $ E_{3} $ may be a nonlinear centre.

    Remark that, as the parameters move away from $ H, $ towards the region containing the limit cycle generated by the Hopf bifurcation, the cycle may disappear, either by transforming into a homoclinic loop or by becoming too large, and no longer inside the visible neighborhood of the origin. In such cases there should exist a bifurcation curve $ L $ originating at $ \mu = 0, $ along which those situations may occur. The position of the curve $ L $ depends on the type of the Hopf bifurcation. These situations also appear in the "difficult case" in the nondegenerate double-Hopf bifurcation. The existence and location of the curve $ L $ can be proved following the lines in [1,2].

    Theorem 4.7. For all $ \gamma > 0, N > 0, $ and $ (\theta _{2}, \delta) $ in regionsVI$ _{\theta } $, VII$ _{\theta } $, VIII$ _{\theta } $ in the $ \left(\theta_{2}, \delta \right) - $ plane, the bifurcation curves consist of

    $ \begin{equation*} O\cup T_{1}\cup T_{3}\cup \Delta _{\pm }\cup X_{-}\cup X_{+}\cup Y_{-}^{\prime }\cup Y_{+}^{\prime }\cup H\cup L. \end{equation*} $

    The generic parameter portraits are shown in Figure 7, in the hypothesis that the Lyapunov coefficient is negative. The corresponding generic phase portraits are given in Figure 8. Many of the phase portraits in Figures 6 and 8 cannot be found in the nondegenerate double-Hopf bifurcation.

    The effect of adding higher-order terms to the truncated normal form system (2.16) and the correspondence between the properties of the 2D amplitude system (2.16) and the 4D system (2.1), have been described, for example, in [1,2,15,16,17,18]. However, the study of nonsymmetric general perturbations of the truncated normal forms is far from complete, as pointed out in [1].

    Thus, we remark that the study of adding higher order terms to the truncated normal form is a complex open problem which is outside the aim of this article.

    The classical normal form of the Hopf-Hopf bifurcation, in differential systems of dimension four and having minimum two independent parameters, is based on six generic conditions (2.8).

    In this work we have studied two degenerate Hopf-Hopf bifurcations, namely:

    1) the case when the generic condition (HH.3) in system (2.8) is not fulfilled, treated in Section 3;

    2) the case when the generic condition (HH.1) in system (2.8) is not fulfilled, treated in Section 4.

    In our study we have obtained new bifurcation diagrams and new phase portraits which were not previously reported in generic studies on double-Hopf bifurcation. Such an example is a saddle-node bifurcation in the amplitude system which corresponds to a fold bifurcation of cycles in the four dimensional system and which is present only in the degenerate case $ \theta = 0 $.

    The results describing the dynamics, contribute to a better understanding of the behavior of a system presenting a Hopf-Hopf singularity.

    Finally, we summarize in the Table 1 the correspondence of equilibria, cycles and bifurcations between the 2D amplitude system (2.16) and the 4D normal form system (2.14):

    Table 1.  The correspondence of equilibria, cycles and bifurcations.
    2D 4D
    $ E_{0} $ origin
    $ E_{3} $ 2D torus
    cycle 3D torus
    $ X_{+}, X_{-}, Y_{+}, Y_{-}, Y_{+}^{\prime }, Y_{-}^{\prime } $ Hopf bifurcation
    $ T_{1}, T_{2}, T_{3} $ Neimark-Sacker bifurcation of cycles
    $ \Delta _{+}, \Delta _{-} $ fold bifurcation of cycles
    Hopf bifurcation $ H $ branching of 3D torus from a 2D torus
    cycle blow up $ L $ blow-up of a 3D torus

     | Show Table
    DownLoad: CSV

    This research was supported by Horizon2020-2017-RISE-777911 project.

    The authors declare there is no conflicts of interest.



    [1] S. M. Allen, J. W. Cahn, A microscopic theory for antiphase boundary motion and its application to antiphase domain coarsening, Acta Metall., 27 (1979), 1085-1095.
    [2] A. A. Wheeler, W. J. Boettinger, G. B. McFadden, Phase-field model for isothermal phase transitions in binary alloys, Phys. Rev. A., 45 (1992), 7424-7439.
    [3] M. Beneŝ, V. Chalupecky, K. Mikula, Geometrical image segmentation by the Allen-Cahn equation, Appl. Numer. Math., 51 (2004), 187-205.
    [4] X. B. Feng, A. Prohl, Numerical analysis of the Allen-Cahn equation and approximation for mean curvature flaws, Numer. Math., 94 (2003), 33-65.
    [5] D. S. Cohen, J. D. Murray, A generalized diffusion model for growth and dispersal in a population, J. Math. Biol., 12 (1981), 237-249.
    [6] L. Q. Chen, Phase-field models for microstructure evolution, Ann. Rev. Mater. Res., 32 (2002), 113-140.
    [7] X. F. Chen, C. Elliott, A. Gardiner, J. J. Zhao, Convergence of numerical solutions to the Allen-Cahn equation, Appl. Anal., 69 (1998), 47-56.
    [8] S. Y. Zhai, X. L. Feng, Y. N. He, Numerical simulation of the three dimensional Allen-Cahn equation by the high-order compact ADI method, Comput. Phys. Commun., 185 (2014), 2449-2455.
    [9] T. L. Hou, T. Tang, J. Yang, Numerical analysis of fully discretized Crank-Nicolson scheme for fractional-in-space Allen-Cahn equations, J. Sci. Comput., 72 (2017), 1214-1231.
    [10] X. B. Feng, Y. K. Li, Analysis of symmetric interior penalty discontinuous Galerkin methods for the Allen-Cahn equation and the mean curvature flow, IMA J. Numer. Anal., 35 (2014), 1622-1651.
    [11] C. Y. Li, Y. Q. Huang, N. Y. Yi, An unconditionally energy stable second order finite element method for solving the Allen-Cahn equation, J. Comput. Appl. Math., 353 (2019), 38-48.
    [12] Z. F. Weng, L. K. Tang, Analysis of the operator splitting scheme for the Allen-Cahn equation, Numer. Heat. TR. B-Fund., 70 (2016), 472-483.
    [13] D. Jeong, J. Kim, An explicit hybrid finite difference scheme for the Allen-Cahn equation, J. Comput. Appl. Math., 340 (2018), 247-255.
    [14] Y. Q. Huang, W. Yang, H. Wang, J. T. Cui, Adaptive operator splitting finite element method for Allen-Cahn equation, Numer. Methods Partial Differ. Eq., 35 (2019), 1290-1300.
    [15] J. Berrut, L. N. Trefethen, Barycentric Lagrange interpolation, SIAM Rev., 46 (2004), 501-517.
    [16] J. P. Berrut, G. Klein, Recent advances in linear barycentric rational interpolation, J. Comput. Appl. Math., 259 (2014), 95-107.
    [17] H. Y. Liu, J. Huang, Y. B. Pan, J. Zhang, Barycentric interpolation collocation methods for solving linear and nonlinear high-dimensional fredholm integral equations, J. Comput. Appl. Math., 327 (2018), 141-154.
    [18] M. Li, C. M. Huang, W. Y. Ming, Barycentric rational collocation methods for Volterra integral equations with weakly singular kernels, Comput. Appl. Math., 38 (2019), 1-15.
    [19] W. H. Luo, T. Z. Huang, X. M. Gu, Y. Liu, Barycentric rational collocation methods for a class of nonlinear parabolic partial differential equations, Appl. Math. Lett., 68 (2017), 13-19.
    [20] O. Oruc, Two meshless methods based on local radial basis function and barycentric rational interpolation for solving 2D viscoelastic wave equation, Comput. Math. Appl., 79 (2020), 3272-3288.
    [21] S. C. Yi, L. Q. Yao, A steady barycentric Lagrange interpolation method for the 2D higher-order time-fractional telegraph equation with nonlocal boundary condition with error analysis, Numer. Methods Partial Differ. Eq., 35 (2019), 1694-1716.
  • This article has been cited by:

    1. Juanjuan Huang, Qinsheng Bi, Mixed-mode bursting oscillations in the neighborhood of a triple Hopf bifurcation point induced by parametric low-frequency excitation, 2023, 166, 09600779, 113016, 10.1016/j.chaos.2022.113016
    2. Raluca Efrem, Mihaela Sterpu, Local dynamics and bifurcation for a two-dimensional cubic Lotka-Volterra system (I), 2023, 50, 12236934, 247, 10.52846/ami.v50i1.1779
    3. Sterpu Mihaela, Efrem Raluca , Local dynamics and bifurcation for a two-dimensional cubic Lotka-Volterra system (Part II), 2024, 51, 12236934, 561, 10.52846/ami.v51i2.2070
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(3090) PDF downloads(195) Cited by(12)

Figures and Tables

Figures(6)  /  Tables(3)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog