Research article Special Issues

Summer cooling potential of urban vegetation—a modeling study for Melbourne, Australia

  • The summer cooling potential of urban vegetation is investigated using an urban climate model for the current and future climates in the Melbourne central business district (CBD) area with various urban forms and vegetation schemes. Simulation results suggest that the average seasonal summer temperatures can be reduced in the range of around 0.5 and 2°C if the Melbourne CBD were replaced by vegetated suburbs and planted parklands, respectively, benefiting a reduction in the number of hot days. It was also found that despite the projected warming in the future and variations in the climate projections among different climate models, the average seasonal cooling potential due to various urban vegetation schemes may not change significantly in comparison with those predicted for the current climate, indicating little dependency on climate change. This finding suggests that the average seasonal cooling potential as a result of urban vegetation in future climates may be empirically quantified in similar amounts to those under the current climate. When urban climate models are used, the cooling potential of urban vegetation in future climates may be quantified by modeling several selected years with one or a few climate models.

    Citation: Dong Chen, Marcus Thatcher, Xiaoming Wang, Guy Barnett, Anthony Kachenko, Robert Prince. Summer cooling potential of urban vegetation—a modeling study for Melbourne, Australia[J]. AIMS Environmental Science, 2015, 2(3): 648-667. doi: 10.3934/environsci.2015.3.648

    Related Papers:

    [1] Azzh Saad Alshehry, Humaira Yasmin, Rasool Shah, Roman Ullah, Asfandyar Khan . Numerical simulation and analysis of fractional-order Phi-Four equation. AIMS Mathematics, 2023, 8(11): 27175-27199. doi: 10.3934/math.20231390
    [2] Mamta Kapoor, Nehad Ali Shah, Wajaree Weera . Analytical solution of time-fractional Schrödinger equations via Shehu Adomian Decomposition Method. AIMS Mathematics, 2022, 7(10): 19562-19596. doi: 10.3934/math.20221074
    [3] Rasool Shah, Abd-Allah Hyder, Naveed Iqbal, Thongchai Botmart . Fractional view evaluation system of Schrödinger-KdV equation by a comparative analysis. AIMS Mathematics, 2022, 7(11): 19846-19864. doi: 10.3934/math.20221087
    [4] Qasem M. Tawhari . Advanced analytical techniques for fractional Schrödinger and Korteweg-de Vries equations. AIMS Mathematics, 2025, 10(5): 11708-11731. doi: 10.3934/math.2025530
    [5] Maysaa Al-Qurashi, Saima Rashid, Fahd Jarad, Madeeha Tahir, Abdullah M. Alsharif . New computations for the two-mode version of the fractional Zakharov-Kuznetsov model in plasma fluid by means of the Shehu decomposition method. AIMS Mathematics, 2022, 7(2): 2044-2060. doi: 10.3934/math.2022117
    [6] Rehana Ashraf, Saima Rashid, Fahd Jarad, Ali Althobaiti . Numerical solutions of fuzzy equal width models via generalized fuzzy fractional derivative operators. AIMS Mathematics, 2022, 7(2): 2695-2728. doi: 10.3934/math.2022152
    [7] Aslı Alkan, Halil Anaç . A new study on the Newell-Whitehead-Segel equation with Caputo-Fabrizio fractional derivative. AIMS Mathematics, 2024, 9(10): 27979-27997. doi: 10.3934/math.20241358
    [8] Mubashir Qayyum, Efaza Ahmad, Hijaz Ahmad, Bandar Almohsen . New solutions of time-space fractional coupled Schrödinger systems. AIMS Mathematics, 2023, 8(11): 27033-27051. doi: 10.3934/math.20231383
    [9] Fouad Mohammad Salama, Nur Nadiah Abd Hamid, Norhashidah Hj. Mohd Ali, Umair Ali . An efficient modified hybrid explicit group iterative method for the time-fractional diffusion equation in two space dimensions. AIMS Mathematics, 2022, 7(2): 2370-2392. doi: 10.3934/math.2022134
    [10] Aslı Alkan, Halil Anaç . The novel numerical solutions for time-fractional Fornberg-Whitham equation by using fractional natural transform decomposition method. AIMS Mathematics, 2024, 9(9): 25333-25359. doi: 10.3934/math.20241237
  • The summer cooling potential of urban vegetation is investigated using an urban climate model for the current and future climates in the Melbourne central business district (CBD) area with various urban forms and vegetation schemes. Simulation results suggest that the average seasonal summer temperatures can be reduced in the range of around 0.5 and 2°C if the Melbourne CBD were replaced by vegetated suburbs and planted parklands, respectively, benefiting a reduction in the number of hot days. It was also found that despite the projected warming in the future and variations in the climate projections among different climate models, the average seasonal cooling potential due to various urban vegetation schemes may not change significantly in comparison with those predicted for the current climate, indicating little dependency on climate change. This finding suggests that the average seasonal cooling potential as a result of urban vegetation in future climates may be empirically quantified in similar amounts to those under the current climate. When urban climate models are used, the cooling potential of urban vegetation in future climates may be quantified by modeling several selected years with one or a few climate models.


    For decades, many researchers have studied and established mathematical models describing interactions among species. Initially, most studies primarily focused on various types of interaction mechanisms and models using random dispersal of species. The importance of species dispersal, which is one of the main factors involved in species behavior, led to the establishment of more realistic models for species dispersal, for example, self-cross diffusion [1,2,3,4,5,6,7,8,9,10], advection-diffusion [11,12,13,14,15,16,17,18,19,20], and other types of non-uniform dispersal [21,22,23,24,25,26,27,28] models. For the predator–prey interaction, a popular way of describing the dispersal mechanism of predators is using prey-taxis [29,30,31,32,33,34]. There are some predator dispersal types such as the cross diffusion [2,3,4,5,6] and starvation-driven diffusion [22,25]. However, most models describing the dispersal mechanism of predators were dependent on prey's population density.

    In this paper, we examine the predator–prey model describing a situation wherein the predator dispersal depends on other environmental factors rather than prey density. Here, we consider the following model with a directional dispersal of a predator in a spatially heterogeneous environment:

    $ \left\{ ut=μΔu+u(m(x)u)α(x)uvu+d(x)vvt=(ηvχvβ)+v(c+β(x)uu+d(x)v)inΩ×(0,),un=ηvnχvβn=0onΩ×(0,),u(x,0)=u0(x)0,v(x,0)=v0(x)0inΩ,
    \right. $
    (P)

    where $ \Omega $ is a bounded convex domain in $ \mathbb{R}^n $. The zero-flux condition is provided on the smooth boundary $ \partial\Omega $, where $ \vec{n} $ is the outward unit normal vector on the boundary. The functions $ u(x, t) $ and $ v(x, t) $ represent the population densities of the prey and predator, respectively. $ m(x) $ is the growth rate of the prey depending on the location $ x\in\Omega $, and $ \mu $ and $ \eta $ are the diffusion rates of the prey and predator, respectively. $ \alpha(x) $ is a positive function representing the capturing rate. $ \beta(x) $ is proportional to $ \alpha(x) $, $ \beta = k\alpha $, where $ k > 0 $ is a conversion efficiency of food into offspring. A positive function $ d(x) $ is a measure of the interference for the predators during hunting the prey. $ c $ and $ \chi $ are positive constants representing the death rate of the predator and taxis sensitivity coefficient, respectively. The reaction term of the predator represents a case wherein the predator is fed only using the prey. To describe the predator–prey interaction, a ratio-dependent functional response is adopted. By considering the spatially heterogeneous environment, we use the predator-dependent form, such as the ratio-dependent form, for the system [35,36,37].

    The model (P) represents a situation wherein the predator dispersal is unaffected by the prey density. The evolution of prey's behavior and traits has helped them against the threats of the predators. Consequently, these evolved traits and behaviors, for example, nocturnality [38,39], camouflage and masquerade [40,41], or living underground [42], make it difficult for the predators to chase the prey directly. This implies that the prey density does not affect the predator dispersal anymore. In response to defensive mechanisms of the prey, predators not only memorize the location where the prey was detected before but also learn the locations to hunt the prey in their habitat efficiently. Some studies have investigated the searching mechanism depending on such spatial memory [43,44,45,46]. In [43], the artificial experiments using birds was introduced, which provided some evidence of such animal learning to find better locations (see [43] and references therein). Thus, we assume that the prey has evolved to avoid detection from the predator, and the predator disperses using a spatial memory from experiences about the good locations to forage efficiently.

    Such predator dispersal is mathematically described using the taxis term $ -\nabla\cdot(\chi v\nabla \beta) $. An attractant function $ \beta(x) $ includes factors such as an encounter with prey and attack success rates. That is, $ \beta(x) $ indicates which area has high or low hunting efficiency for predators. Thus, the taxis term represents that the predator moves toward locations with high foraging efficiency using the spatial memory about $ \beta $, which is a kind of mnemotaxis [44,45]. The reaction–diffusion system with the taxis term has a property that $ v $ is concentrated around the local maximums of $ \beta $ [15,16]. This implies that the predator gathers near the locations that have a high hunting efficiency. Thus, the predators' directional dispersal toward the location with better hunting efficiency can be described using such type of taxis term.

    For the described dispersal mechanism of the predators, two situations can occur. The first situation wherein $ \beta $ is proportional to the resources of prey $ m $ ($ \beta = \kappa m $ for some constant $ \kappa $). This means that the predators move toward locations where the prey is abundant because the prey is distributed along with the distribution of resources. Mathematical and experimental works have been performed on a similar situation that the predators track the prey's resources instead of the prey density [47,48,49]. The second situation in which the distributions of $ \beta $ and $ m $ are different, i.e., $ \beta $ is not proportional to $ m $. This situation can happen because some places are rich with prey but are easy to hide, making it difficult to hunt for predators, or because the predators memorized correctly, but the environment of prey's habitat has changed due to external factors. This situation seems to be disadvantageous to the predators. Thus, this paper aimed to show the effect of the directional movement of predators in both situations.

    In our model, we let $ w = e^{ -\chi \beta (x)/\eta }v $. Then, an alternative form of (P) is obtained as follows:

    $ \left\{ ut=μΔu+u(m(x)u)α(x)uwu+d(x)weχβ(x)eχβ(x)/ηwt=ηΔw+χβw+w(c+β(x)uu+d(x)weχβ(x)/η)inΩ×(0,),un=wn=0onΩ×(0,),u(x,0)=u0(x)0,w(x,0)=w0(x)0inΩ,
    \right. $
    (P')

    where $ w(x_0) = e^{ -\chi \beta (x)/\eta }v_0 $. Throughout this paper, we give assumptions on $ m $, $ \alpha $ and $ \beta $:

    $ m(x),\alpha (x),\beta (x)\;\;{\rm{and}}\;\;d(x)\;\;{\rm{are}}\;\;{\rm{in}}\;\;{C^2}(\bar \Omega ),\;\;\;\int_\Omega {m(x)dx} \gt 0 $ (A)

    Under the assumption (A), the global existence and uniqueness of the classical solution to (P') follows from the Amann's results [50] for quasilinear parabolic equations. And (P') has one semi-trivial solution $ (\theta, 0) $, where $ \theta $ is a unique positive steady state solution of

    $ {μΔu+u(m(x)u)=0inΩ,un=0onΩ.
    $
    (1.1)

    In this paper, we first investigate the local stability of the semi-trivial solution $ (\theta, 0) $, which is closely related to the survival of the predator. We also consider the elliptic system of (P'). We show that the elliptic system has a positive solution, called the coexistence steady-state, under the conditions for the instability of $ (\theta, 0) $. We also investigate the nonexistence of the positive solution.

    The remaining part of this paper is organized as follows. In Section 2, we state the main theorems and give a biological explanation. The theorems address the stability of $ (\theta, 0) $ and the existence of a coexistence steady-state. Moreover, the effects of taxis sensitivity $ \chi $ are observed. We provide the proofs of stability theorem using the principal eigenvalue analysis in Section 3. The existence of a coexistence steady-state is shown by the fixed-point index theory in Section 4. In Section 5, we present the numerical simulations for some cases. Section 6 summarizes the results obtained and concludes the paper.

    First, we introduce the results for the stability of $ (\theta, 0) $. The instability of $ (\theta, 0) $ and the positivity of the prey density imply that the predator can invade the region. This implies that the predator can survive in the environment, even though this is rare. This allows us to examine whether the taxis is beneficial to the survival of the predator. As in [29], we can show the positivity of the prey density and $ \liminf_{t\rightarrow\infty}u(x, t) > 0 $ if $ \int_\Omega{m(x)dx} > \int_\Omega{\alpha(x)dx} $. Therefore, if we assume $ \int_\Omega{m(x)dx} > \int_\Omega{\alpha(x)dx} $, the invasibility of the predator can be determined by the instability of $ (\theta, 0) $.

    To study the stability of $ (\theta, 0) $, we consider the following linearized eigenvalue problem of (P') at $ (\theta, 0) $:

    $ {μΔϕ+(m(x)2θ)ϕα(x)eχβ(x)/ηψ=λϕinΩ,ηΔψ+χβψ+(c+β(x))ψ=λψinΩ,ϕn=ψn=0onΩ.
    $
    (2.1)

    We multiply $ e^{\chi\beta/\eta} $ by the second equation of Eq (2.1) and obtain

    $ η(eχβ/ηψ)+(c+β(x))eχβ/ηψ=λeχβ/ηψ.
    $
    (2.2)

    Equation (2.2) is the weighted eigenvalue problem $ \lambda e^{\chi\beta/\eta}\psi = L\psi $, where $ L\psi = \eta\nabla\cdot(e^{\chi\beta/\eta}\nabla\psi) + (-c+\beta(x))e^{\chi\beta/\eta}\psi $. Since $ L $ is a self-adjoint, we have the following variational form of the principal eigenvalue:

    $ λ1=supψH1(Ω){0},Ωeχβ/ηψ2=1Ωηeχβ/η|ψ|2+(c+β(x))eχβ/ηψ2.
    $

    We denote the principal eigenvalue of the operator $ L $ by $ \lambda_1(L, \chi) $. It is well-known that $ \lambda_1(L, \chi) $ and the local stability of $ (\theta, 0) $ have the following relation: If $ \lambda_1(L, \chi) > 0 $, then $ (\theta, 0) $ is locally unstable; if $ \lambda_1(L, \chi) < 0 $, then $ (\theta, 0) $ is locally stable (see [51]).

    We define the ecological reproduction number $ \mathcal{R}_0 $ as

    $ R0=supψH1(Ω){0}Ωβ(x)eχβ/ηψ2Ωηeχβ/η|ψ|2+ceχβ/ηψ2>0.
    $
    (2.3)

    Thus, we have the following lemma whose proof is analogous to the proof of Lemma 2, 3 in [52].

    Lemma 2.1. $ \mathcal{R}_0 > 1 $ if and only if $ \lambda_1(L, \chi) > 0 $.

    Lemma 2.1 implies that $ (\theta, 0) $ is unstable if $ \mathcal{R}_0 > 1 $, and $ (\theta, 0) $ is locally asymptotically stable if $ \mathcal{R}_0 < 1 $. Then, $ \mathcal{R}_0 $ has a similar role as the basic reproduction number in epidemiology. This means that the survival of the predator is determined by the sign of $ \mathcal{R}_0-1 $.

    Now, we introduce the results for the local stability of $ (\theta, 0) $ with an assumption on $ \beta $, $ \frac{\partial\beta}{\partial \vec{n}} = 0 $. For a function $ f $, we denote the average of $ f $ in $ \Omega $ by $ \overline{f} $. We first present the stability result about the predator's death rate $ c $ when $ \frac{\partial\beta}{\partial \vec{n}} = 0 $.

    Theorem 2.2. Let $ \eta > 0, \chi\geq0 $ be given. If $ \frac{\partial\beta}{\partial \vec{n}} = 0 $, there exists $ \tilde{c}\in \big(\overline{\beta}-\frac{\chi^2}{4\eta}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $ such that $ (\theta, 0) $ is unstable when $ c < \tilde{c} $, and $ (\theta, 0) $ is locally asymptotically stable when $ c > \tilde{c} $.

    In Theorem 2.2, the sharp criteria for the local stability of $ (\theta, 0) $ about $ c $ is obtained, which represents invasibility of the predator. Given $ \eta > 0 $ and $ \chi\geq0 $, there exists a threshold value $ \tilde{c} $ such that $ \lambda_1(L, \chi) = 0 $ for $ c = \tilde{c} $ and the sign of $ c-\tilde{c} $ determines the stability of $ (\theta, 0) $. From Theorem 2.2, $ (\theta, 0) $ is unstable for all $ \eta > 0 $ when $ c < \overline{\beta} $ and $ \chi = 0 $, but $ \chi > 0 $ gives a possibility $ (\theta, 0) $ to be stable if $ c\in(\overline{\beta}-\frac{\chi^2}{4\eta}\overline{|\nabla\beta|}, \overline{\beta}) $. This means that the directional dispersal could be disadvantageous.

    Next, we give the result for the local stability of $ (\theta, 0) $ about the death rate $ c $ and diffusion rate $ \eta $ of a predator.

    Theorem 2.3. Suppose that $ \frac{\partial\beta}{\partial \vec{n}} = 0 $. Let $ \chi > 0 $ and $ \xi > 0 $ be given constants.

    (i) If $ c > \max_{\bar{\Omega}}\beta $, then $ (\theta, 0) $ is locally asymptotically stable for all $ \eta > 0 $.

    (ii) If $ c < \min_{\bar{\Omega}}\beta $, then $ (\theta, 0) $ is unstable for all $ \eta > 0 $.

    (iii) If $ c < \overline{\beta}-\frac{1}{4\xi}\overline{|\nabla\beta|^2} $, then $ (\theta, 0) $ is unstable when $ \eta > \xi\chi^2 $.

    (iv) Suppose that $ c\in \big(\overline{\beta}-\frac{1}{4\xi}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $. Then $ \tilde{\eta} > 0 $ exists such that if $ \tilde{\eta} < \xi\chi^2 $ then $ (\theta, 0) $ is locally asymptotically stable for $ \eta\in(\tilde{\eta}, \xi\chi^2) $; and if $ \tilde{\eta} > \xi\chi^2 $ then $ (\theta, 0) $ is unstable for $ \eta\in(\xi\chi^2, \tilde{\eta} $).

    Theorem 2.3 gives sufficient conditions of the stability of $ (\theta, 0) $ for the predators' diffusion rate $ \eta $ when $ c $ is in a proper region. A positively given $ \xi $ plays a role for obtaining the sufficient conditions for the stability. For example, we suppose that $ c $ and $ \chi $ such that $ c < \overline{\beta}-\frac{1}{4\xi_1}\overline{|\nabla\beta|^2} $ and $ c\in \big(\overline{\beta}-\frac{1}{4\xi_2}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $ for some $ \xi_1 > \xi_2 $. Then, we obtain from Theorem 2.3 that $ (\theta, 0) $ is unstable for $ \eta > \xi_1\chi^2 $. Moreover, $ (\theta, 0) $ is unstable for $ \eta\in(\xi_2\chi^2, \tilde{\eta}(\xi_2)) $ if $ \tilde{\eta(\xi_2)} > \xi_2\chi^2 $, and $ (\theta, 0) $ is locally asymptotically stable for $ \eta\in(\tilde{\eta}, \xi_2\chi^2) $ if $ \tilde{\eta} < \xi_2\chi^2 $.

    Now, we investigate the effect of taxis $ \chi $ for fixed $ c $ and $ \eta $.

    Theorem 2.4. Suppose that $ \beta $ has at least one isolated global maximum point and $ \max_{\bar{\Omega}}\beta > c $. For given $ \eta > 0 $, there exists $ \tilde{\chi} > 0 $ such that if $ \chi\geq \tilde{\chi} $ then $ (\theta, 0) $ is unstable.

    From Theorem 2.4, we can compare the stability of $ (\theta, 0) $ with or without the predator's directional movement. Without the taxis ($ \chi = 0 $), the predator cannot survive when they are rare for $ (c, \eta)\in (\overline{\beta}, \max_{\bar{\Omega}}\beta)\times (\eta^*, \infty) $ for some $ \eta^* $, which will be immediately shown by lemma in Section 3. However, Theorem 2.4 shows that a large taxis $ \chi > 0 $ makes the predators invade a region. This implies that the strong taxis for predators gives a survival advantage to them compared with a random dispersal.

    Next, we give the final result for the stability of $ (\theta, 0) $ when $ \beta(x) = \kappa m(x) $ for $ x\in\Omega $. Consider the following equation $ A\chi^2 + B\chi + C = 0 $, where

    $ A=κ24ηΩ|m|2θ2dx,B=κΩθθmdx,C=Ωη|θ|2+(κmc)θ2dx.
    $
    (2.4)

    If $ B $ is positive and $ B^2-4AC > 0 $, the equation has two real roots $ \chi_1 < \chi_2 $ such that $ \chi_2 $ is positive.

    Theorem 2.5. Suppose that $ m $ is positive, $ \beta = \kappa m $ for some constant $ \kappa > 0 $ and $ \max_{\bar{\Omega}}\beta > c $. If $ B^2-4AC > 0 $, $ (\theta, 0) $ is unstable when $ \chi\in(\max\{\chi_1, 0\}, \chi_2) $.

    Remark that $ \beta(x) = \kappa m(x) $ implies that the predator moves toward the favorable habitat of the prey. Because the prey has a property to be distributed according to $ m(x) $, such movement of the predator allows them to go to an environment where the prey population is sufficient. Moreover, as the assumption $ \beta(x) = \kappa m(x) $ gives two effects, which are taxis toward the location with an abundance of prey and high hunting efficiency, the fitness of the predator increases. Therefore, a proper taxis sensitivity provides a survival advantage to the predator without the assumption on $ \beta $ in Theorem 2.4 for which predators cannot survive without taxis (Theorem 2.5).

    Consider the steady-state of the system (P'):

    $ {μΔu+u(m(x)u)α(x)uwu+d(x)weχβ(x)/ηeχβ(x)/η=0ηΔw+χβw+w(c+β(x)uu+d(x)weχβ(x)/η)=0inΩ,un=wn=0onΩ.
    $
    (2.5)

    In Eq (2.5), the ratio-dependent functional response is not defined at $ (u, w) = (0, 0) $. Since

    $ \lim\limits_{(u, w)\rightarrow (0, 0)} \frac{uwe^{ \chi \beta (x)/\eta }}{ u + d(x)we^{ \chi \beta (x)/\eta }} = 0, $

    we can extend the domain of $ \frac{ud(x)we^{ \chi \beta (x)/\eta }}{u + d(x)we^{ \chi \beta (x)/\eta }} $ to $ \{(u, w) : u\geq 0, w\geq 0\} $, so that $ (0, 0) $ becomes a trivial solution of Eq (2.5).

    Let a positive solution of Eq (2.5) be a coexistence steady-state. Then, we have the following results for the existence of the coexistence steady-state of Eq (2.5), which is proven using the fixed-point index theory.

    Theorem 2.6. Let $ \eta > 0 $ and $ \chi\geq0 $ be given. If $ \frac{\partial\beta}{\partial \vec{n}} = 0 $, there exists $ \tilde{c}\in \big(\overline{\beta}-\frac{\chi^2}{4\eta}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $ such that Eq (2.5) has at least one coexistence steady state if and only if $ c < \tilde{c} $.

    Theorem 2.7. Suppose that $ \frac{\partial\beta}{\partial \vec{n}} = 0 $. Let $ \chi > 0 $ and $ \xi > 0 $ be given constants.

    (i) If $ c < \overline{\beta}-\frac{1}{4\xi}\overline{|\nabla\beta|^2} $, then Eq (2.5) has at least one coexistence steady state when $ \eta > \xi\chi^2 $.

    (ii) If $ c < \min_{\bar{\Omega}}\beta $, then Eq (2.5) has at least one coexistence steady state for all $ \eta > 0 $.

    (iii) Suppose that $ c\in \big(\overline{\beta}-\frac{1}{4\xi}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $. Then $ \tilde{\eta} > 0 $ exists such that if $ \tilde{\eta} < \xi\chi^2 $ then Eq (2.5) has at least one coexistence steady state for $ \eta\in(\xi\chi^2, \tilde{\eta} $).

    Theorem 2.8. Suppose that $ \beta $ has at least one isolated global maximum point and $ \max_{\bar{\Omega}}\beta > c $. For given $ \eta > 0 $, there exists $ \tilde{\chi} > 0 $ such that if $ \chi\geq \tilde{\chi} $ then Eq (2.5) has at least one coexistence steady state.

    Theorem 2.9. Suppose that $ m $ is positive, $ \beta = \kappa m $ for some constant $ \kappa > 0 $ and $ \max_{\bar{\Omega}}\beta > c $. If $ B^2-4AC > 0 $ for given $ A, B $ and $ C $ as Eq (2.4), then Eq (2.5) has at least one coexistence steady state when $ \chi\in(\max\{\chi_1, 0\}, \chi_2) $.

    It is noteworthy that the conditions for the existence of coexistence steady-state are closely related to the sign of $ \lambda_1(L, \chi) $, which will result in the proof of the above theorems. From the results in subsection 2.1, Eq (2.5) has more chance to have the coexistence steady-state when taxis sensitivity $ \chi $ is properly chosen.

    In this section, we prove Theorem 2.2–2.5. Recall that the principal eigenvalue of (2.2) is defined by

    $ λ1(L,χ)=supψH1(Ω){0},Ωeχβ/ηψ2=1Ωηeχβ/η|ψ|2+(c+β(x))eχβ/ηψ2.
    $
    (3.1)

    We assume that $ \frac{\partial\beta}{\partial\vec{n}} = 0 $ and consider $ \varphi = e^{\chi\beta/2\eta}\psi $. Since $ \psi\in H^1(\Omega) $ and $ \beta\in C^2(\bar{\Omega}) $, $ \varphi $ is also in $ H^1(\Omega) $. In addition, a mapping $ \psi\mapsto e^{\chi\beta/2\eta}\psi $ is bijective. Thus, if we take $ \psi = e^{-\chi\beta/2\eta}\varphi $ and take supremum over $ \varphi\in H^1(\Omega) $, we can rewrite the (3.1) as

    $ λ1(L,χ)=supφH1(Ω){0},||φ||2=1Ωη|φ|2+χφφβ+(χ24η|β|2+β(x)c)φ2.
    $
    (EV1)

    Then, by Green's identity and the assumption of $ \beta $, we have

    $ λ1(L,χ)=supφH1(Ω){0},||φ||2=1Ωη|φ|2+(χ2Δβχ24η|β|2+β(x)c)φ2.
    $
    (EV2)

    It can be easily shown from (EV2) that $ \lambda_1(L, \chi) $ is the principal eigenvalue of

    $ {ηΔφ+(χ2Δβχ24η|β|2+β(x)c)=λφinΩ,φn=0onΩ.
    $

    Let $ \psi_1 $ be the principal eigenfunction with respect to $ \lambda_1(L, \chi) $. From (EV2), we obtain

    $ \lambda_1(L, \chi) = \int_\Omega{-\eta|\nabla\varphi_1|^2 + \Big(-\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big)\varphi_1^2} , $

    where $ \varphi_1 = e^{\chi\beta/2\eta}\psi_1 $. Since the principal eigenfunction is unique, $ \varphi_1 $ is the principal eigenfunction with respect to $ \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) $. Therefore, $ \lambda_1(L, \chi) = \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) $. We will utilize the forms (EV1) and (EV2) in proving our results.

    Before proving the theorems, we introduce some properties of the principal eigenvalue. The first property is well-known; the proof is omitted here (See [53]).

    Lemma 3.1. Let $ \lambda_1(\eta\Delta+g(x)) $ be the principal eigenvalue of

    $ {ηΔψ+g(x)ψ=λψinΩ,ψn=0onΩ.
    $
    (3.2)

    Then,

    (i) $ \lambda_1(\eta\Delta+g(x)) $ is monotone decreasing function to $ \eta $;

    (ii) $ \lambda_1(\eta\Delta+\tilde{g}(x))\geq\lambda_1(\eta\Delta+g(x)) $ for a function $ \tilde{g}(x)\geq g(x) $;

    (iii) $ \lim_{\eta\rightarrow0}\lambda_1(\eta\Delta+g(x)) = \max_{x\in\bar{\Omega}}g(x) $;

    (iv) $ \lim_{\eta\rightarrow\infty}\lambda_1(\eta\Delta+g(x)) = \overline{g}. $

    Lemma 3.2. Let $ \lambda_1(\eta\Delta+g(x)) $ be the principal eigenvalue of (3.2).

    (i) If $ g(x) < 0 $, then $ \lambda_1(\eta\Delta+g(x)) < 0 $.

    (ii) If $ \overline{g} > 0 $, then $ \lambda_1(\eta\Delta+g(x)) > 0 $.

    (iii) If $ \int_\Omega{g(x)dx} < 0 $ and $ g(x) $ is positive somewhere, then there exists $ \tilde{\eta} > 0 $ such that $ \lambda_1(\eta\Delta+g(x))\geq0 $ if and only if $ \eta\leq\tilde{\eta} $. The equality holds when $ \eta = \tilde{\eta} $.

    Proof. (i) is obvious by the definition of the principal eigenvalue.

    (ii) Suppose $ \int_\Omega{g(x)dx} > 0 $. From the definition of $ \lambda_1(\eta\Delta + g(x)) $,

    $ λ1(ηΔ+g(x))=supψH1(Ω){0},||ψ||2=1Ωη|ψ|2+g(x)ψ2dx¯g>0.
    $

    (iii) Let $ S: = \{x:g(x) > 0\}\neq \emptyset $. Then, we can choose a test function $ \tilde{\psi} $ with $ ||\tilde{\psi}||_2 = 1 $ whose support is in $ S $. Then, we obtain

    $ λ1Ωη|˜ψ|2+g(x)˜ψ2dx=Sη|˜ψ|2+g(x)˜ψ2dx>0
    $

    for sufficiently small $ \eta $. Since $ \lambda_1(\eta\Delta+g(x)) $ is monotone decreasing function to $ \eta $ and $ \lim_{\eta\rightarrow\infty}\lambda_1(\eta\Delta+g(x)) = \overline{g} < 0 $, there exists $ \tilde{\eta} $ such that if $ \eta > \tilde{\eta} $, then $ \lambda_1(\eta\Delta+g(x)) < 0 $, and if $ \eta < \tilde{\eta} $, then $ \lambda_1(\eta\Delta+g(x)) > 0 $, which is a desired result.

    Now, we first give the proof Theorems 2.2 and 2.3.

    Proof of Theorem 2.2. From the definition of the principal eigenvalue (3.1), we have

    $ \lambda_1(L, \chi) = \sup\limits_{\psi\in H^1(\Omega) \setminus\{0\}, \int_\Omega{e^{\chi\beta/\eta}\psi^2} = 1}\int_\Omega{-\eta e^{\chi\beta/\eta} |\nabla\psi|^2 + (-c+\beta(x))e^{\chi\beta/\eta}\psi^2} \lt 0 $

    for $ c > \max_{\bar{\Omega}}\beta $.

    Next, from (EV2) and $ \frac{\partial\beta}{\partial\vec{n}} = 0 $, we obtain

    $ λ1(L,χ)=λ1(ηΔχ2Δβχ24η|β|2+(β(x)c))=Ωη|φ1|2+(χ2Δβχ24η|β|2+(β(x)c))φ21,
    $

    where $ \varphi_1 $ is the principal eigenfunction with respect to $ \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) $. Also, by the Green's first identity, we have

    $ ΩΔβ(x)dx=ΩβndS=0.
    $
    (3.3)

    It follows from Eq (3.3) and Lemma 3.2 that

    $ λ1(ηΔχ2Δβχ24η|β|2+(β(x)c))>0forc<¯βχ24η¯|β|2.
    $

    Also, since Eq (3.1) and (EV2) are equivalent form of the principal eigenvalue,

    $ \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) \lt 0\quad\mbox{for}\;\; c \gt \max\limits_{\bar{\Omega}}\beta. $

    By Lemma 3.1 (ii), $ \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) $ is monotone decreasing function to $ c $. Then, we can find $ \tilde{c}\in \big(\overline{\beta} -\frac{\chi^2}{4\eta}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $ such that $ \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) = 0 $. By the monotonicity of $ \lambda_1(L, \chi) $ to $ c $, $ (\theta, 0) $ is unstable when $ c < \tilde{c} $, and $ (\theta, 0) $ is locally asymptotically stable when $ c > \tilde{c} $.

    Proof of Theorem 2.3. As in the proof of Theorem 2.2, it suffices to check the sign of $ \lambda_1(L, \chi) $. From the form Eq (3.1), (i) and (ii) are immediately obtained.

    For (iii), we denote $ \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{\chi^2}{4\eta}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) $ by $ \tilde{\lambda} $. Consider an eigenvalue problem as follows:

    $ {ηΔφ+(χ2Δβ14ξ|β|2+(β(x)c))φ=λφinΩ,φn=0onΩ.
    $
    (3.4)

    Then, the principal eigenvalue of Eq (3.4) can be written by

    $ λ1(ηΔχ2Δβ14ξ|β|2+(β(x)c))=supφH1(Ω){0},||φ||2=1Ωη|φ|2+(χ2Δβ14ξ|β|2+(β(x)c))φ2,
    $

    From Lemma 3.1 (ii),

    $ ˜λ>λ1(ηΔχ2Δβ14ξ|β|2+(β(x)c))if and only if η>ξχ2.
    $
    (3.5)

    Then, the results (iii) follows from Lemma 3.2.

    (iv) Assume that $ c\in \big(\overline{\beta}-\frac{1}{4\xi}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $. By Lemma 3.2 (iii), there exists $ \tilde{\eta} > 0 $ depending on $ \xi $ and $ \chi $ such that $ \lambda_1\Big(\eta\Delta -\frac{\chi}{2} \Delta \beta -\frac{1}{4\xi}|\nabla\beta|^2 +\big(\beta(x)-c\big)\Big) < 0 $ if and only if $ \eta > \tilde{\eta} $. For $ \tilde{\eta} < \xi\chi^2 $, $ (\tilde{\eta}, \xi\chi^2) $ is a non-empty interval. Then, it follows form (3.5) that $ \tilde{\lambda} < 0 $ when $ \eta\in(\tilde{\eta}, \xi\chi^2) $, and thus $ (\theta, 0) $ is locally asymptotically stable. Similarly, for $ \tilde{\eta} > \xi\chi^2 $, $ (\theta, 0) $ is unstable when $ \eta\in(\xi\chi^2, \tilde{\eta}) $.

    Next, we prove Theorem 2.4.

    Proof of Theorem 2.4. It can be proved similarly to Theorem 4.4 in [15]. For reader's convenience, we give the proof here. Since the principal eigenvalue of (2.2) is

    $ \lambda_1(L, \chi) = \sup\limits_{\psi\in H^1(\Omega) \setminus\{0\}, \int_\Omega{e^{\chi\beta/\eta}\psi^2} = 1}\int_\Omega{-\eta e^{\chi\beta/\eta} |\nabla\psi|^2 + (-c+\beta(x))e^{\chi\beta/\eta}\psi^2}, $

    we only need to find a function $ \Psi $ with $ \int_\Omega{e^{\chi\beta/\eta}\Psi^2} = 1 $ such that

    $ Ωηeχβ/η|Ψ|2<Ω(c+β(x))eχβ/ηΨ2.
    $
    (3.6)

    From the assumption, a point $ x_0\in\bar{\Omega} $ exists satisfying $ \beta(x_0) = \max_{\bar{\Omega}}\beta $ and $ \beta(x_0)-c\geq \delta $ for some $ \delta > 0 $. Let $ R_1 > 0 $ be small such that $ \beta(x)-c\geq \frac{1}{2}\delta $ in $ B_{R_1}(x_0)\cap\Omega $, where $ B_{R_1}(x_0) $ is a ball of radius $ R_1 $ and centered at $ x_0 $. We define $ \beta_1 = \max_{(B_{R_1}(x_0)\setminus B_{R_1/2}(x_0))\cap\Omega}\beta $ and $ \beta_2 = \min_{B_{R_2}(x_0)\cap\Omega}\beta $ for $ R_1, R_2 > 0 $. Since the global maximum point is isolated,

    $ \beta_2\geq \frac{1}{2}[\beta_1+\beta(x_0)] \gt \beta_1 $

    for sufficiently small $ R_2 $ satisfying $ R_2\leq R_1/2 $.

    Define a function $ \Phi\in C^1(\bar{\Omega}) $ as

    $ Φ={1in BR1/2(x0)Ω,[0,1]in (BR1(x0)BR2(x0))Ω,0otherwise.
    $

    If we put $ \Psi: = \Phi/\big(\int_\Omega{e^{\chi\beta/\eta}\Phi^2}\big)^{1/2} $ in Eq (3.6), the left-hand side becomes

    $ Ωηeχβ/η|Ψ|2=η(BR1(x0)BR2(x0))Ωeχβ/η|Ψ|2C1(BR1(x0)BR2(x0))Ωeχβ/ηC1|Ω|e(χ/η)β1,
    $

    where $ C_1 $ satisfies $ ||\nabla\Psi||_\infty\leq C_1 $. The right-hand side of Eq (3.6) is

    $ Ω(c+β(x))eχβ/ηΨ2=BR1(x0)Ω(c+β(x))eχβ/ηΨ2e(χ/η)β2BR2(x0)Ω(c+β(x))C2e(χ/η)β2
    $

    for some positive constant $ C_2 $. Since $ C_1 $ and $ C_2 $ is independent of $ \chi $, Eq (3.6) holds if we choose $ \chi $ large enough. Hence, we have a desired result.

    Finally, we assume that $ \beta(x) = \kappa m(x) $ for $ x\in\Omega $, which is the assumption in Theorem 2.5. Before proving the theorem, we give a known result in [14].

    Lemma 3.3 (Cantrell et al. [14]). The integral $ \int_\Omega{\theta \nabla\theta\cdot\nabla m dx} $ is positive if $ \Omega\subseteq\mathbb{R} $ is an interval or if $ \Omega\subseteq \mathbb{R}^n $ is convex.

    Then, we give the proof of Theorem 2.5.

    Proof of Theorem 2.5. Let $ \lambda_1(L, \chi) $ be the principal eigenvalue with the form (EV1). Since $ \beta(x) = \kappa m(x) $ for $ x\in\Omega $,

    $ λ1(L,χ)=supφH1(Ω){0},||φ||2=1Ωη|φ|2+κχφφm+(κ2χ24η|m|2+κmc)φ2Ωη|θ|2+κχθθm+(κ2χ24η|m|2+κmc)θ2/Ωθ2dx=(Aχ2+Bχ+C)/Ωθ2dx,
    $

    where $ A, B $ and $ C $ are given as Equation (2.4). Since $ B^2-4AC > 0 $, $ A < 0 $ and $ B > 0 $ by Lemma 3.3, the equation has two real roots $ \chi_1 $ and $ \chi_2 $ with at least one positive solution, say $ \chi_2 $. Then, $ A\chi^2 +B\chi +C > 0 $ for $ \chi_1 < \chi < \chi_2 $. which implies $ (\theta, 0) $ is unstable for $ \chi\in(\max\{\chi_1, 0\}, \chi_2) $. We note that $ B^2-4AC $ cannot be positive for $ c > \max_{\bar{\Omega}}\beta $, because $ (\theta, 0) $ is locally asymptotically stable ($ \lambda_1 < 0 $) when $ c > \max_{\bar{\Omega}}\beta $. Hence, Theorem 2.5 is valid for $ c < \max_{\bar{\Omega}}\beta $.

    In this section, we present the proofs of Theorems 2.6–2.8 which are about the existence and non-existence of the coexistence steady-states of Eq (2.5). Before proving the theorems, we introduce the fixed-point index theory.

    Fixed-point index theory

    Let $ E $ be a real Banach space and $ W \subset E $ a closed convex set. $ W $ is called a total wedge if $ \alpha W \subset W $ for all $ \alpha \geq 0 $ and $ \overline{W\cup (-W)} = E $. A wedge is said to be a cone if $ W \cap (-W) = \{0\} $. For $ y \in W $, we define $ W_y = \{ x \in E\; :\; y + \gamma x \in W \; \; \mbox{for some}\; \; \gamma > 0 \} $ and $ S_y = \{ x \in {\overline W}_y \; :\; -x \in {\overline W}_y\} $. Then, $ {\overline W}_y $ is a wedge containing $ W $, $ y $ and $ -y $, while $ S_y $ is a closed subspace of $ E $ containing $ y $.

    Let $ T $ be a compact linear operator on $ E $ satisfying $ T({\overline W}_y) \subset {\overline W}_y $. We say that $ T $ has property $ \alpha $ on $ {\overline W}_y $ if there are $ t \in (0, 1) $ and $ w \in {\overline W}_y \backslash S_y $ such that $ w - tTw \in S_y $. Let $ F : W \rightarrow W $ is a compact operator with a fixed-point $ y \in W $ and $ F $ is Fréchet differentiable at $ y $. Let $ L = F'(y) $ be the Fréchet derivative of $ F $ at $ y $. Then, $ L $ maps $ {\overline W}_y $ into itself. For an open subset $ U \subset W $, define $ \operatorname{index}_W(F, U) = \operatorname{index}(F, U, W) = \operatorname{deg}_W (I-F, U, 0) $, where $ I $ is the identity map. If $ y $ is an isolated fixed-point of $ F $, then the fixed-point index of $ F $ at $ y $ in $ W $ is defined by $ \operatorname{index}_W(F, y) = \operatorname{index}(F, y, W) = \operatorname{index}(F, U(y), W) $, where $ U(y) $ is a small open neighborhood of $ y $ in $ W $.

    The following theorem can be obtained from the results of [54,55,56].

    Theorem 4.1. Assume that $ I - L $ is invertible on $ {\overline W}_y $.

    (i) If $ L $ has property $ \alpha $ on $ { \overline W}_y $, then $ \operatorname{index}_W(F, y) = 0 $.

    (ii) If $ L $ does not have property $ \alpha $ on $ { \overline W}_y $, then $ \operatorname{index}_W(F, y) = (-1)^{\sigma} $, where $ \sigma $ is the sum of multiplicities of all the eigenvalues of $ L $ which are greater than 1.

    Lemma 4.2. Any coexistence state $ (u, w) $ of system (2.5) has an a priori estimate

    $ u(x) \leq Q_{1}, \; w(x) \leq Q_{2}. $

    where $ Q_{1} = \max_{\bar{\Omega}} m, \; Q_{2} = \dfrac{Q_1\max_{\bar{\Omega}}\beta}{c\min_{\bar{\Omega}}d} $.

    Proof. Let $ (u, w) $ be a positive solution of (2.5). Since $ -\mu \Delta u \leq u\Big(m(x) - u \Big) $, we obtain $ u \leq \max_{\bar{\Omega}}{m} $ by maximum principle for elliptic.

    Let $ w(x_0) = \max_{\bar{\Omega}} w $ and $ h = \chi / \eta $. Then, by maximum principle for elliptic equation,

    $ 0 \leq \dfrac{ \beta(x_0) u(x_0)}{u(x_0) + d(x_0) e^{h\beta (x_0)} w(x_0)} - c \leq \dfrac{ \max\limits_{\bar{\Omega}} \beta Q_1}{Q_1 + \min\limits_{\bar{\Omega}}d w(x_0)} - c. $

    Since

    $ c \min\limits_{\bar{\Omega}}d w(x_0) \leq cQ_1 + c\min\limits_{\bar{\Omega}}d w(x_0) \leq \max\limits_{\bar{\Omega}}\beta Q_1, $

    it implies that

    $ \max\limits_{\bar{\Omega}}w \leq \frac{\max\limits_{\bar{\Omega}}\beta }{c\min\limits_{\bar{\Omega}}d} Q_1. $

    Notation 4.3.

    (i) $ X : = C^{1}_{N} (\bar{\Omega}) \bigoplus C^{1}_{N} (\bar{\Omega}) $ where $ C^{1}_{N} (\bar{\Omega}): = \{ \phi \in C^{1}(\bar{\Omega}) : \frac{\partial\phi}{\partial \vec{n}} = 0 \; \; on\; \; \partial \Omega \} $.

    (ii) $ Q : = \max\{ Q_1, Q_2 \} + 1 $.

    (iii) $ D : = D_Q \bigoplus D_Q $ where $ D_Q = \{ \phi \in C^{1}_{N}(\bar{\Omega}) : \phi < Q \; \; on\; \; \bar{\Omega} \} $.

    (iv) $ W : = D_0 \bigoplus D_0 $ where $ D_0 = \{ \phi \in C^{1}_{N}(\bar{\Omega}) : 0 \leq \phi \; \; on\; \; \bar{\Omega} \} $.

    (v) $ D^{\prime} : = D \cap W $.

    For $ \tau \in [0, 1] $, define a positive compact operator $ F_{\tau} : X \to X $ by $ F_{\tau}(u, w) = (-D\Delta + P)^{-1}[P\mathcal{I} + \tau G](u, w)^{T} $ where

    $G(u, w)^{T} = \left(muu2αueχβ/ηwu+d(x)eχβ/ηwχβw+βueχβ/ηwu+d(x)eχβ/ηwcw
    \right), \qquad D = (μ00η)
    . $

    and $ P $ is a positive constant such that $ (2Q + \frac{\max_{\Omega}\alpha}{\min_{\Omega}d} + \frac{\max_{\Omega} \beta}{{\min_{\Omega}d}} + c)Q + \chi \Vert \nabla \beta \Vert_{\infty}\Vert \nabla w \Vert_{\infty} < P $. $ F_{\tau} $ is positive and compact operator. We denote $ F = F_1 $. Then, (P') has a positive solution if and only if $ F $ has a positive fixed-point.

    For the operator $ F $, we first find the result for the index values, $ \operatorname{index}_{W}(F, D^{\prime}) $ and $ \operatorname{index}_{W}(F, (0, 0)) $. The following lemmas can be proved by similar argument to [57]; the proofs are omitted here.

    Lemma 4.4. $ \operatorname{index}_{W}(F, D^{\prime}) = 1 $.

    Lemma 4.5. $ \operatorname{index}_{W}(F, (0, 0)) = 0 $.

    Next, we calculate the $ \operatorname{index}_{W}(F, (\theta, 0)) $. Remark that the principal eigenvalue $ \lambda_1(L, \chi) $ of Eq (2.2) is real, and the corresponding principal eigenfunction $ \Psi $ is positive. Since the principal eigenpair $ (\lambda_1(L, \chi), \Psi) $ satisfies Eq (2.2), $ (\lambda_1(L, \chi), \Psi) $ also satisfies the second equation of Eq (2.1):

    $ λ1(L,χ)Ψ=ηΔΨ+χβΨ+(c+β(x))ΨinΩ.
    $
    (4.1)

    Lemma 4.6. If $ \lambda_{1}(L, \chi) > 0 $, then $ \operatorname{index}_{W}(F, (\theta, 0)) = 0 $.

    Proof. By calculation, we have

    $ \overline{W}_{(\theta, 0)} = C^{1}_{N}( \overline{\Omega}) \bigoplus D_0 , \;\;S_{(\theta, 0)} = C^{1}_{N}( \overline{\Omega}) \bigoplus \{0 \}. $

    Define

    $ L:=F(θ,0)=(DΔ+P)1(m2θ+Pα0χβ+βc+P).
    $

    We first claim that $ \mathcal{I}-\mathcal{L} $ is invertible. Let $ \mathcal{L}(\phi, \psi)^{T} = (\phi, \psi)^{T} \in \overline{W}_{(\theta, 0)} $.

    $ {μΔϕ=(m2θ)ϕαψηΔψ=χβψ+(βc)ψinΩ,ϕn=ψn=0onΩ.
    $
    (4.2)

    For $ \psi \geq 0 $, multiply the second equation of Eq (4.2) by $ e^{\chi\beta / \eta}\Psi $ and integrate over $ \Omega $, where $ \Psi $ is a principal eigenfunction of $ \lambda_{1}(L, \chi) $. Then, we obtain

    $ 0=Ωeχβ/ηΨ[ηΔψ+χβψ]+eχβ/η(βc)Ψψ=ΩΨ[ηeχβ/ηψ]+eχβ/η(βc)Ψψ=Ωψ[ηeχβ/ηΨ]+eχβ/η(βc)Ψψ=Ωeχβ/ηψ[ηΔΨ+χβΨ+(βc)Ψ]=λ1(L,χ)Ωeχβ/ηΨψ0,
    $

    which implies $ \psi\equiv0 $. The last equality holds from Eq (4.1).

    Since $ \lambda_1(\mu\Delta + (m-\theta)) = 0 $ with the corresponding principal eigenfunction $ \theta $, $ \lambda_1(\mu\Delta + (m-2\theta)) < 0 $ by Lemma 3.1 (ii). We multiply the first equation of Eq (4.2) by $ \phi $ and integrate over $ \Omega $. Then, we have

    $ 0 = \int_{\Omega} \alpha \phi \psi = \int_{\Omega}-\mu |\nabla \phi|^2 + (m - 2\theta) \phi^2 \leq \lambda_1(\mu\Delta + (m-2\theta))\int_{\Omega} {\phi^2} \leq 0. $

    Hence, $ \phi \equiv 0 $ and $ \mathcal{I}-\mathcal{L} $ is invertible.

    Next, we show that $ \mathcal{L} $ has a property $ \alpha $. Let $ r $ be $ \mbox{the spectral radius of} \; (-\eta \Delta + P)^{-1}(\chi \nabla \beta \cdot \nabla + \beta - c + P) $. Then, the assumption $ \lambda_{1}(L, \chi) > 0 $ implies that $ r > 1 $. Then, there exist corresponding eigenfunction $ \xi $ exists in $ D_0\setminus\{0\} $. Take $ t = 1 /r \in (0, 1) $, then

    $ \left(ϕψ
    \right) - t \mathcal{L} \left(ϕψ
    \right) = \left(00
    \right) \in S_{(\theta, 0)} \;\;{\rm{holds }}\;\;{\rm{for}} \quad \left(ϕψ
    \right) = \left(0ξ
    \right) \in \overline{W}_{(\theta, 0)}\setminus S_{(\theta, 0)}.$

    Hence, $ \mathcal{L} $ has property $ \alpha $. By Theorem 4.1 (i), we have $ \operatorname{index}_{W}(F, (\theta, 0)) = 0 $.

    Now, we give the proofs of Theorems 2.6–2.9. Theorems 2.7–2.9 can be proved in a similarly manner; we only give the proof of Theorem 2.6.

    Proof of Theorem 2.6. Suppose that $ \frac{\partial\beta}{\partial \vec{n}} = 0 $. From Theorem 2.2, there exists $ \tilde{c}\in \big(\overline{\beta}-\frac{\chi^2}{4\eta}\overline{|\nabla\beta|^2}, \max_{\bar{\Omega}}\beta \big) $. Then, it follows from Lemma 4.4–4.6 that

    $ 1 = \operatorname{index}_{W}(F, D') \neq \operatorname{index}_{W}(F, (0, 0)) + \operatorname{index}_{W}(F, (\theta, 0)) = 0. $

    Hence, Eq (2.5) has at least one coexistence state for $ c < \tilde{c} $.

    Next, we assume that $ c\geq \tilde{c} $. Suppose that $ (u, w) $ be the positive solution of Eq (2.5). Then, $ (u, w) $ satisfies

    $ ηΔw+χβw+w(c+β(x)uu+d(x)weχβ(x)/η)=0.
    $
    (4.3)

    Let $ \tilde{w}: = w\big/\big(\int_\Omega{e^{\chi\beta/\eta}w^2}\big)^{1/2} $. If we multiply Eq (4.3) by $ e^{\chi\beta/\eta}\tilde{w}/\big(\int_\Omega{e^{\chi\beta/\eta}w^2}\big)^{1/2} $ and integrate over $ \Omega $, we have

    $ 0=Ωηeχβ/η|˜w|2+(β(x)uu+d(x)weχβ/ηc)eχβ/η˜w2=Ωηeχβ/η|˜w|2+(β(x)c)eχβ/η˜w2+(β(x)uu+d(x)weχβ/ηβ(x))eχβ/η˜w2λ1(L,χ)+Ω(β(x)uu+d(x)weχβ/ηβ(x))eχβ/η˜w2.
    $
    (4.4)

    The last inequality holds by the definition of the principal eigenvalue $ \lambda_1(L, \chi) $. From Eq (4.4), we obtain

    $ \lambda_1(L, \chi)\geq \int_\Omega{\Big(\beta(x)-\frac{\beta(x)u}{u+d(x)we^{\chi\beta/\eta}}\Big) e^{\chi\beta/\eta}\tilde{w}^2} \gt 0, $

    which is a contradiction because $ \lambda_1(L, \chi)\leq0 $ under the assumption $ c\geq\tilde{c} $. Hence, there is no positive solution of Eq (2.5) if $ c\geq \tilde{c} $.

    In this section, we present some numerical simulations to verify the mathematical results of this paper. Throughout the simulations, the following parameters and functions are used:

    $ β(x)=0.8α(x),d(x)=0.5+0.1sin(2πx),μ=0.01,u0(x)=0.5+0.2sin(2πx),v0(x)=0.2+0.1sin(2πx),Ω=[0,1].
    $

    We used the upwind scheme with the finite difference method, and the toleration of iteration is selected as $ 10^{-6}. $ The simulation results presented show the important role of directional dispersal with taxis compared to the random dispersal when $ c, \beta $ and $ \eta $ are given.

    Figure 1 shows that a large $ \chi $ gives the predator with directional dispersal a survival advantage when $ c\in\big(\overline{\beta}, \max_{\bar{\Omega}}\beta\big) $. Theorem 2.4 implies that the $ (\theta, 0) $ is locally asymptotically stable when $ \chi = 0 $ for any given $ c $ and $ \eta $, but a large $ \chi $ allows $ (\theta, 0) $ to be unstable. This means that the strong taxis of predator can make them survive under the condition in which the predator with random dispersal cannot survive. Figure 1 shows that such dispersal makes the predator and prey coexist in the environment. It is noteworthy that, as can be seen in Figure 1, the density of the predator is concentrated on the local maximum of $ \beta $, which is a feature of the species with directional dispersal.

    Figure 1.  Stability of semi-trivial solution $ (\theta, 0) $ when $ \overline{\beta} < c < \max_{\bar{\Omega}}\beta $. (a) $ (\theta, 0) $ is locally asymptotically stable when $ \chi = 0 $. (b) $ (\theta, 0) $ is unstable when $ \chi = 0.3 $ $ (\beta = 0.48-0.32\cos(4\pi x), m(x) = 0.6 + 0.4\sin(3\pi x), c = 0.7, \eta = 0.01) $.

    In Figure 2, the case that $ \beta(x) = \kappa m(x) $ is considered. If $ \beta(x) $ has at least one isolated global maximum point, we have a same result in Theorem 2.4 and Figure 1. Thus, we give $ \beta(x) $ as a combination of smoothstep function for Figure 2, which is the case when $ \beta $ has no isolated global maximum point on $ \overline{\Omega} $. Figure 2 shows that the properly chosen taxis sensitivity $ \chi > 0 $ can make the predator survive under the condition when the predator without taxis cannot survive. This represents the result of Theorem 2.5 that the directional dispersal is beneficial for the predator's survival when $ \chi $ is in a proper region regardless of the assumption on $ \beta $. This is a natural ecological result because the taxis of the predators not only makes them move towards the region with high hunting efficiency, but also the prey's favorable habitats where the density of the prey is large.

    Figure 2.  Stability of semi-trivial solution $ (\theta, 0) $ when $ \beta(x) = \kappa m(x) $. (a) $ (\theta, 0) $ is locally asymptotically stable when $ \chi = 0 $. (b) $ (\theta, 0) $ is unstable when $ \chi = 0.01 $. $ (\beta(x) = 0.72m(x), c = 0.73, \eta = 0.01) $.

    In this paper, we suggested the predator–prey model describing the phenomena that the predator moves with taxis toward the region with a high hunting efficiency. For the predator–prey model, most of the studies dealing with taxis have considered the prey-taxis [29,30,31,32,33,34]. Authors in [29] adopted the ratio-dependent functional responses with a spatially heterogeneous environment, which are the same as our model, and they showed that the prey taxis is always beneficial for the predator's survival. However, the predators cannot chase the prey directly because the model we established describes the situation that prey has evolved to avoid detection from the predator. Thus, we considered other type of predators' dispersal mechanism with taxis toward the location where the hunting efficiency is high using the spatial memory [43,44,45,46]. There are two cases of the rate at which predators capture the prey: proportional ($ \beta = \kappa m $) or not proportional ($ \beta\neq \kappa m $) to the resources of prey. For the first case, a similar idea is introduced in [47,48,49] that the predators move in the direction of high prey's resources instead of tracking the prey directly, and [47] established the mathematical model of such dispersal and obtained some mathematical results. The other case can be followed from the prey's defensive mechanisms or environmental change from other external factors. This could make the predators go to a region with low prey density, which seems not helpful for the predators. In this paper, we investigated whether the defensive mechanism of the predators is advantageous to them for both cases.

    Through this article, we determined the criteria and sufficient conditions for the local stability of the semi-trivial solution, which is closely related to the survival of the predators. Furthermore, the conditions for the existence of coexistence steady-state were obtained. We studied the stability and coexistence conditions for the three parameters: the death rate, diffusion, and taxis sensitivity. Finally, the obtained results were verified by using numerical simulations. From a biological perspective, the results for stability indicate that, when the predators have the spatial memory where the location is good to forage efficiently, the taxis toward the location could give a greater chance of invasion than when they disperse randomly (Theorems 2.2, 2.4 and 2.5). This implies that the directional movement toward the region with high foraging efficiency increases the fitness of predators. Consequently, such dispersal helps the predators to invade a region. If $ \beta $ has at least one isolated global maximum point, the dispersal is advantageous to the predator's survival when taxis is strong even though $ \beta $ is not proportional to $ m $ (Theorem 2.4). This represents that the dispersal mechanism also affects to the predators fitness regardless the distributions of $ \beta $ and $ m $. When $ \beta $ has no isolated global maximum, the proper taxis sensitivity gives survival advantage to the predators (Theorem 2.5). This means that the properly chosen taxis sensitivity helps predators survive regardless of assumption on $ \beta $ when the predators have directional movement toward the location with abundant prey and high hunting efficiency ($ \beta $ is proportional to $ m $). Moreover, as a dictum in ecology that "invasibility implies coexistence, " the taxis influences on the coexistence of predator and prey. This dispersal of predators could give them a greater possibility to coexist in the ecosystem than random dispersal (Theorems 2.6, 2.8 and 2.9). Thus, we conclude that the dispersal considered in this work plays a vital role in the survival of the predator and the coexistence in the predator–prey system.

    This research was supported by Basic Science Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, ICT & Future Planning (NRF-2015R1A2A2A01007013; NRF-2018R1C1B5039938).

    All authors declare no conflicts of interest in this paper.

    [1] United Nations, World Urbanization Prospects, the 2014 Revision. 2014. Available from: http://esa.un.org/unpd/ accessed March 2015.
    [2] DSEWPaC (Department of Sustainability, Environment, Water, Population and Communities), State of the Environment 2011, 2011. Available from: http://www.environment.gov.au/science/soe/2011-report/contents
    [3] Morris CJG, Simmonds I, Plummer N (2001) Quantification of the Influences of Wind and Cloud on the Nocturnal Urban Heat Island of a Large City. J Appl Meteorol 40: 169-182. doi: 10.1175/1520-0450(2001)040<0169:QOTIOW>2.0.CO;2
    [4] Coutts A, Beringer J, Tapper N (2010) Changing urban climate and CO2 emissions: implications for the development of policies for sustainable cities. Urban Policy Res 28: 27-47. doi: 10.1080/08111140903437716
    [5] Wu K, Yang XQ (2013) Urbanization and heterogeneous surface warming in eastern China. Chin Sci Bull 58: 1363-1373. doi: 10.1007/s11434-012-5627-8
    [6] DSE (Department of Sustainability and Environment), Melbourne 2030: Planning for Sustainable Growth. State Government of Victoria, 2002. Available from: http://www.dtpli.vic.gov.au/planning/plans-and-policies/planning-for-melbourne/melbournes-strategic-planning-history/melbourne-2030-planning-for-sustainable-growth
    [7] Konopacki S, Akbari H (2002) Energy savings for heat island reduction strategies in Chicago and Houston (including updates for Baton Rouge, Sacramento, and Salt Lake City). Draft Final Report, LBNL-49638, University of California, Berkeley.
    [8] Kolokotroni M, Giannitsaris I, Watkins R (2006) The effect of the London Urban Heat Island on building summer cooling demand and night ventilation strategies. Sol Energ 80: 383-392. doi: 10.1016/j.solener.2005.03.010
    [9] Priyadarsini R (2011) Urban Heat Island and its Impact on Building Energy Consumption. Adv Build Energ Res 3: 261-270.
    [10] Changnon SA, Kunkel KE, Reinke BC (1996) Impacts and responses to the 1995 heat wave: A call to action. Bull Am Meteorol Soc 77: 1497-1505. doi: 10.1175/1520-0477(1996)077<1497:IARTTH>2.0.CO;2
    [11] Haines A, Kovats RS, Campbell-Lendrum D, et al. (2006) Climate change and human health: impacts, vulnerability, and mitigation. Lancet 367: 2101-2109. doi: 10.1016/S0140-6736(06)68933-2
    [12] Cadot E, Rodwin VG, Spira A (2007) In the heat of the summer: lessons from the heat waves in Paris. J Urban Health 84: 466-468. doi: 10.1007/s11524-007-9161-y
    [13] Luber G (2008) Climate Change and Extreme Heat Events, Am J Prev Med 35: 429-435.
    [14] Nicholls N, Skinner C, Loughnan M, et al. (2008) A simple heat alert system for Melbourne, Australia, Int J Biometeorol 52: 375-384.
    [15] Kovats RS, Hajat S (2008) Heat stress and public health: a critical review. Annu Rev Public Health 29: 41-55. doi: 10.1146/annurev.publhealth.29.020907.090843
    [16] Loughnan ME, Nicholls N, Tapper NJ (2010) The effects of summer temperature, age and socioeconomic circumstance on acute myocardial infarction admissions in Melbourne, Australia. Int J Health Geogr 9: 41. doi: 10.1186/1476-072X-9-41
    [17] DHS, January 2009 Heatwave in Victoria: an assessment of health impacts. Victorian Government Department of Human Services, Victoria, 2009. Available from: http://docs.health.vic.gov.au/docs/doc/F7EEA4050981101ACA257AD80074AE8B/$FILE/heat_health_impact_rpt_Vic2009.pdf
    [18] Alexander LV, Arblaster J (2008) Assessing trends in observed and modelled climate extremes over Australia in relation to future projections. Int J Climatol 29: 417-435.
    [19] Solecki WD, Rosenzweig C, Parshall L, et al. (2005) Mitigation of the heat island effect in urban New Jersey. Global Environ Chang Part B: Environ Hazards 6: 39-49.
    [20] Gill SE, Handley JF, Ennos AR, et al. (2008) Characterising the urban environment of UK cities and towns: A template for landscape planning. Landscape Urban Plann 87: 210-222. doi: 10.1016/j.landurbplan.2008.06.008
    [21] Gaitani N, Spanou A, Saliari M, et al. (2011) Improving the microclimate in urban areas: a case study in the centre of Athens. Build Serv Eng Res Technol 32: 53-71. doi: 10.1177/0143624410394518
    [22] Kleerekoper L, van Esch M, Salcedo TB (2012) How to make a city climate proof, addressing the urban heat island effect. Resour Conserv Recycl 64: 30-38. doi: 10.1016/j.resconrec.2011.06.004
    [23] DCLG, Building a greener future: towards zero carbon development. UK: Department of Communities and Local Government, 2007. Available from: http://webarchive.nationalarchives.gov.uk/20120919132719/http://www.communities.gov.uk/documents/planningandbuilding/pdf/153125.pdf
    [24] European Commission, Low energy buildings in Europe: current state of play, definition and best practice. 2009. Available from:http: //regions202020.eu/cms/inspiration/resources/PublicResource/330/view
    [25] Wang XM, Chen D, Ren ZG (2010) Assessment of climate change impact on residential building heating and cooling energy requirement in Australia. Build Environ 45: 1663-1682. doi: 10.1016/j.buildenv.2010.01.022
    [26] Rosenfeld AH, Akbari H, Romm JJ, et al. (1998) Cool communities: strategies for heat island mitigation and smog reduction. Energ Build 28: 51-62. doi: 10.1016/S0378-7788(97)00063-7
    [27] Akbari H, Pomerantz M, Taha H (2001) Cool surfaces and shade trees to reduce energy use and improve air quality in urban areas. Sol Energ 70: 295-310. doi: 10.1016/S0038-092X(00)00089-X
    [28] U.S. Environmental Protection Agency, Reducing Urban Heat Islands: Compendium of Strategies. 2008. Available from: http: //www.epa.gov/hiri/resources/compendium.htm
    [29] Memon RA, Leung DYC, Liu CH (2008) A review on the generation, determination and mitigation of Urban Heat Island. J Environ Sci 20: 120-128. doi: 10.1016/S1001-0742(08)60019-4
    [30] Santamouris M (2012) Cooling the cities - a review of reflective and green roof mitigation technologies to fight heat island and improve comfort in urban environments. Sol Energ 103: 680-703.
    [31] Liu K, Bass B (2005) Performance of Green Roof Systems. National Research Council Canada, Report No. NRCC-47705, Toronto, Canada.
    [32] Rosenzweig C, Solecki W, Parshall L, et al. (2006) Mitigating New York City's Heat Island with Urban Forestry, Living Roofs, and Light Surfaces. Sixth Symposium on the Urban Environment and Forum on Managing our Physical and Natural Resources, American Meteorological Society, Atlanta, GA.
    [33] Alexandri E, Jones P (2008) Temperature decreases in an urban canyon due to green walls and green roofs in diverse climates. Build Environ 43: 480-493. doi: 10.1016/j.buildenv.2006.10.055
    [34] Wong JKW, Lau LSK (2013) From the ‘urban heat island’ to the ‘green island’? A preliminary investigation into the potential of retrofitting green roofs in Mongkok district of Hong Kong. Habit Int 39: 25-35.
    [35] Susca T, Gaffin SR, Dell'Osso GR (2011) Positive effects of vegetation: Urban heat island and green roofs, Environ Pollut 159: 2119-2126.
    [36] Xu T, Sathaye J, Akbari H, et al. (2012) Quantifying the direct benefits of cool roofs in an urban setting: Reduced cooling energy use and lowered greenhouse gas emissions. Build Environ 48: 1-6.
    [37] Smith GB, Aguilar JLC, Gentle AR, et al. (2012) Multi-parameter sensitivity analysis: A design methodology applied to energy efficiency in temperate climate houses. Energy Build 55: 668-673. doi: 10.1016/j.enbuild.2012.09.007
    [38] Bozonnet E, Musy M, Calmet I, et al. (2013) Modeling methods to assess urban fluxes and heat island mitigation measures from street to city scale. Int J Low Carbon Technol 0: 1-16.
    [39] Sailor DJ (1995) Simulated urban response to modifications in surface albedo and vegetative cover. J Appl Meteor 34: 1694- 1704. doi: 10.1175/1520-0450-34.7.1694
    [40] Taha H, Douglas S, Haney J (1997) Mesoscale meteorological and air quality impacts of increased urban albedo and vegetation. Energ Build 25: 169-177. doi: 10.1016/S0378-7788(96)01006-7
    [41] Shashua-Bar L, Hoffman ME (2000) Vegetation as a climatic component in the design of an urban street: An empirical model for predicting the cooling effect of urban green areas with trees. Energy Build 31: 221-35. doi: 10.1016/S0378-7788(99)00018-3
    [42] Kong FH, Yin HW, Wang CZ, et al. (2014) A satellite image-based analysis of factors contributing to the green-space cool island intensity on a city scale. Urban For Urban Gree 13: 846-853. doi: 10.1016/j.ufug.2014.09.009
    [43] Perini K, Magliocco A (2014) Effects of vegetation, urban density, building height, and atmospheric conditions on local temperatures and thermal comfort. Urban For Urban Gree 13:495-506. doi: 10.1016/j.ufug.2014.03.003
    [44] Chen C, Yue Y, Jiang W (2014) Numerical Simulation on Cooling Effects of Greening for Alleviating Urban Heat Island Effect in North China. Appl Mech Mater 675-677: 1227-1233. doi: 10.4028/www.scientific.net/AMM.675-677.1227
    [45] Yang JB, Liu HN, Sun JN, et al. (2015) Further Development of the Regional Boundary Layer Model to Study the Impacts of Greenery on the Urban Thermal Environment. J Appl Meteor Climatol 54: 137-152. doi: 10.1175/JAMC-D-14-0057.1
    [46] Gromke C, Blocken B, Janssen W, et al. (2015) CFD analysis of transpirational cooling by vegetation: Case study for specific meteorological conditions during a heat wave in Arnhem, Netherlands, Build Environ 83: 11-26.
    [47] Middel A, Chhetri N, Quay R (2015) Urban forestry and cool roofs: Assessment of heat mitigation strategies in Phoenix residential neighborhoods. Urban For Urban Gree 14: 178-186. doi: 10.1016/j.ufug.2014.09.010
    [48] Thatcher M, Hurley P (2012) Simulating Australian urban climate in a mesoscale atmospheric numerical model. Boundary Layer Meteorol 142: 149-175. doi: 10.1007/s10546-011-9663-8
    [49] Hurley P, Physick W, Luhar A (2005) TAPM - A practical approach to prognostic meteorological and air pollution modeling. Environ Model Sotfw 20: 737-752. doi: 10.1016/j.envsoft.2004.04.006
    [50] Masson V (2000) A physically-based scheme for the urban energy budget in atmospheric models. Boundary Layer Meteorol 94: 357-397. doi: 10.1023/A:1002463829265
    [51] Harman I, Barlow J, Belcher S (2004) Scalar fluxes from urban street canyons. Part II: Model. Boundary Layer Meteorol 113: 387-409. doi: 10.1007/s10546-004-6205-7
    [52] Coutts A, Beringer J, Tapper N (2007) Impact of increasing urban density on local climate: spatial and temporal variations in the surface energy balance in Melbourne, Australia. J Appl Meteor Climatol 46: 477-493. doi: 10.1175/JAM2462.1
    [53] McGregor J, C-CAM: Geometric aspects and dynamical formulation. CSIRO Atmospheric Research, 2005. Available from: www.cmar.csiro.au/e-print/open/mcgregor_2005a.pdf
    [54] McGregor J, Dix M (2008) An updated description of the conformal-cubic atmospheric model. High Resolution Simulation of the Atmosphere and Ocean, Hamilton, K. and Ohfuchi, W., Eds., Springer 51-76.
    [55] Katzfey J, McGregor J, Nguyen K, et al. Dynamical downscaling techniques: impacts on regional climate change signals. In: Anderssen RS, Braddock RD and Newham LTH (ed) 18th World IMACS congress and MODSIM09 international congress on modelling and simulation. 2009, July; Cairns. Modelling and Simulation Society of Australia and New Zealand. pp 2377-2383.
    [56] Nguyen K, Katzfey J, McGregor J (2011) Global 60 km simulations with the Conformal Cubic Atmospheric Model: Evaluation over the tropics. Clim Dyn 39: 637-654.
    [57] City of Melbourne, Urban Forest Strategy Make a Great City Greener 2013-2032. 2015. Available from: http://www.melbourne.vic.gov.au/Sustainability/UrbanForest/Documents/Urban_Forest_Strategy.pdf
  • This article has been cited by:

    1. Qasim Khan, Anthony Suen, Hassan Khan, Poom Kumam, Comparative analysis of fractional dynamical systems with various operators, 2023, 8, 2473-6988, 13943, 10.3934/math.2023714
    2. Muhammad Saqib, Daud Ahmad, Ahmad N. Al-Kenani, Tofigh Allahviranloo, Fourth- and fifth-order iterative schemes for nonlinear equations in coupled systems: A novel Adomian decomposition approach, 2023, 74, 11100168, 751, 10.1016/j.aej.2023.05.047
  • Reader Comments
  • © 2015 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(7464) PDF downloads(1295) Cited by(6)

Figures and Tables

Figures(8)  /  Tables(2)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog