
As a guanylate binding protein (GBPs) member, GBP3 is immune-associated and may participate in oncogenesis and cancer therapy. Since little has been reported on GBP3 in this field, we provide pan-cancer bioinformatics to investigate the role of GBP3 in human cancers. The GBP3 expression, related clinical outcomes, immune infiltrates, potential mechanisms and mutations were conducted using tools including TIMER2.0, GEPIA2.0, SRING, DAVID and cBioPortal. Results showed an increased risk of high GBP3 in Brain Lower Grade Glioma (LGG) and Lung Squamous Cell Carcinoma (LUSC) and a decreased risk of GBP3 in Sarcoma (SARC) and Skin Cutaneous Melanoma (SKCM) (p ≤ 0.05). GBP3 was negatively correlated with CAFs in Esophageal Adenocarcinoma (ESCA) and positively correlated with CAFs in LGG, LUSC and TGCG (p ≤ 0.05). In addition, GBP3 was positively correlated with CD8+ T cells in Bladder Urothelial Carcinoma (BLCA), Cervical Squamous Cell Carcinoma (CESC), Kidney Renal Clear Cell Carcinoma (KIRC), SARC, SKCM, SKCM-Metastasis and Uveal Melanoma (UVM) (p ≤ 0.05). Potentially, GBP3 may participate in the homeostasis between immune and adaptive immunity in cancers. Moreover, the most frequent mutation sites of GBP3 in cancers are R151Q/* and K380N. This study would provide new insight into cancer prognosis and therapy.
Citation: Tongmeng Jiang, Pan Jin, Guoxiu Huang, Shi-Cheng Li. The function of guanylate binding protein 3 (GBP3) in human cancers by pan-cancer bioinformatics[J]. Mathematical Biosciences and Engineering, 2023, 20(5): 9511-9529. doi: 10.3934/mbe.2023418
[1] | Chenxi Huang, Qianqian Zhang, Sanyi Tang . Non-smooth dynamics of a SIR model with nonlinear state-dependent impulsive control. Mathematical Biosciences and Engineering, 2023, 20(10): 18861-18887. doi: 10.3934/mbe.2023835 |
[2] | Roberto A. Saenz, Herbert W. Hethcote . Competing species models with an infectious disease. Mathematical Biosciences and Engineering, 2006, 3(1): 219-235. doi: 10.3934/mbe.2006.3.219 |
[3] | Hui Cao, Dongxue Yan, Ao Li . Dynamic analysis of the recurrent epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 5972-5990. doi: 10.3934/mbe.2019299 |
[4] | Cruz Vargas-De-León, Alberto d'Onofrio . Global stability of infectious disease models with contact rate as a function of prevalence index. Mathematical Biosciences and Engineering, 2017, 14(4): 1019-1033. doi: 10.3934/mbe.2017053 |
[5] | Xin-You Meng, Tao Zhang . The impact of media on the spatiotemporal pattern dynamics of a reaction-diffusion epidemic model. Mathematical Biosciences and Engineering, 2020, 17(4): 4034-4047. doi: 10.3934/mbe.2020223 |
[6] | Zhen Jin, Zhien Ma . The stability of an SIR epidemic model with time delays. Mathematical Biosciences and Engineering, 2006, 3(1): 101-109. doi: 10.3934/mbe.2006.3.101 |
[7] | Peter Rashkov, Ezio Venturino, Maira Aguiar, Nico Stollenwerk, Bob W. Kooi . On the role of vector modeling in a minimalistic epidemic model. Mathematical Biosciences and Engineering, 2019, 16(5): 4314-4338. doi: 10.3934/mbe.2019215 |
[8] | Min Zhu, Xiaofei Guo, Zhigui Lin . The risk index for an SIR epidemic model and spatial spreading of the infectious disease. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1565-1583. doi: 10.3934/mbe.2017081 |
[9] | Ilse Domínguez-Alemán, Itzel Domínguez-Alemán, Juan Carlos Hernández-Gómez, Francisco J. Ariza-Hernández . A predator-prey fractional model with disease in the prey species. Mathematical Biosciences and Engineering, 2024, 21(3): 3713-3741. doi: 10.3934/mbe.2024164 |
[10] | Toshikazu Kuniya, Yoshiaki Muroya, Yoichi Enatsu . Threshold dynamics of an SIR epidemic model with hybrid of multigroup and patch structures. Mathematical Biosciences and Engineering, 2014, 11(6): 1375-1393. doi: 10.3934/mbe.2014.11.1375 |
As a guanylate binding protein (GBPs) member, GBP3 is immune-associated and may participate in oncogenesis and cancer therapy. Since little has been reported on GBP3 in this field, we provide pan-cancer bioinformatics to investigate the role of GBP3 in human cancers. The GBP3 expression, related clinical outcomes, immune infiltrates, potential mechanisms and mutations were conducted using tools including TIMER2.0, GEPIA2.0, SRING, DAVID and cBioPortal. Results showed an increased risk of high GBP3 in Brain Lower Grade Glioma (LGG) and Lung Squamous Cell Carcinoma (LUSC) and a decreased risk of GBP3 in Sarcoma (SARC) and Skin Cutaneous Melanoma (SKCM) (p ≤ 0.05). GBP3 was negatively correlated with CAFs in Esophageal Adenocarcinoma (ESCA) and positively correlated with CAFs in LGG, LUSC and TGCG (p ≤ 0.05). In addition, GBP3 was positively correlated with CD8+ T cells in Bladder Urothelial Carcinoma (BLCA), Cervical Squamous Cell Carcinoma (CESC), Kidney Renal Clear Cell Carcinoma (KIRC), SARC, SKCM, SKCM-Metastasis and Uveal Melanoma (UVM) (p ≤ 0.05). Potentially, GBP3 may participate in the homeostasis between immune and adaptive immunity in cancers. Moreover, the most frequent mutation sites of GBP3 in cancers are R151Q/* and K380N. This study would provide new insight into cancer prognosis and therapy.
Symbiotic relationships play a crucial role in the study of ecological systems and have gained substantial attention among evolutionary biologists for their frequent involvement in the coevolutionary dynamics between interacting organisms. In addition to well-known symbiotic relationships like mutualism, competition and predator/prey interactions, there are three other types, i.e., commensalism, amensalism and synnecrosis, that have received comparatively less attention from researchers. Amensalism, receiving less attention than commensalism, involves one species adversely affecting another's fitness without gaining any apparent benefit [1] For instance, Xi et al. [2] noted that grassland caterpillars will trigger a death-feigning anti-predator response if grasshoppers suddenly appear. Moreover, by leaping to seek food, the grasshopper may negatively affect the foraging efficiency of the caterpillars, but the behavior of caterpillars has no effect on the grasshopper. Therefore, the two species form an amensalism relationship. The interaction of ibex and weevils feeding on the same type of shrub serves as another illustration of amensalism. Although the presence of the weevil has minimal impact on food availability, the presence of ibex significantly reduces weevil populations. Ibex consume substantial amounts of plant matter, inadvertently ingesting the weevils along with it [3]. According to Veiga et al.[4] the relationships witnessed between honey bees (Apis mellifera) and wasps in the tropical ecosystems of East Africa could be described by amensalism concerning nesting spaces. The experiment's results revealed that wasps refrained from using nest boxes that had been inhabited by bees in the prior breeding season. In contrast, honeybees displayed a tendency to take over these nest boxes, regardless of whether they were previously occupied by wasps. This behavior showcases a clear instance of amensalism, as there was no recorded occurrence of wasps occupying a nest box previously utilized by bees. For a deeper ecological context regarding species amensalism, refer to [3,5,6,7,8] and the related references within those sources.
Although scientists and researchers have extensively investigated amensalism in ecological systems in recent decades, the first mathematical model was put forward by Sun in 2003 [9]. In 2008, Zhu et al. [10] employed the method of vector field analysis to study the following model:
{dxdt=x(a1−b1x−c1y),dydt=y(a2−c2y), | (1.1) |
where x and y represent the density of the first and second species at time t, respectively, a1 and b1 stand for the intrinsic growth rates and the intensity of competition of x, a2 and c2 stand for the intrinsic growth rates and the intensity of competition of y. Additionally, c1 represents the strength of the interspecific competition, indicating how much impact the second species has on the first species. By analyzing the equilibrium points in system (1.1), they obtained the following result:
Theorem A (1) If a1c1<a2c2, then the boundary fixed points B(0,a2c2) is globally stable.
(2) If a1c1>a2c2, then the unique positive fixed point A(a1c2−c1a2b1c2,a2c2) is globally stable.
The model (1.1) exhibits simple dynamical behavior as the Theorem A indicates that the system does not generate bifurcation. In recent years, researchers have extensively studied the dynamical behaviors of the amensalism model (1.1) in different scenarios. Several mathematical models have been explored, considering the well-known Allee effect in [11,12,13,14,15]. For instance, Wei et al. [11] proposed an amensalism model with a weak Allee effect and observed saddle-node bifurcations. Luo et al. [12] presented the global dynamics of a two-species Holling-II amensalism system with a nonlinear growth rate by incorporating the Allee effect on the first species. Similarly, some interesting works using non-monotonic functional responses have been presented in [6,16,18]. Amensalism relationships among different species have also been incorporated into the study of harvesting models [19,20]. Under the Michaelis-Menten type harvesting in the first species, Liu et al. [19] investigated the complex dynamics of an amensalism system. Zhao et al. [20] found that incorporating Menten-type harvesting in the second species within the amensalism system leads to notably richer dynamics. These dynamics include scenarios where the extinction of the first species persists or where the approach toward the steady-state occurs at a slower rate. Similar to the above, amensalism models incorporating prey refuge have emerged as a prominent area of research [19,21,22,24].
In ecological systems, direct killing in the predation process significantly influences the common symbiosis relationships, i.e., mutualisms, competition and predator/prey interactions. Besides direct killing, the psychological fear induced in the prey due to predation risk also indirectly impacts the breeding rate of the prey. This indirect effect, termed the fear effect, exerts a substantial impact throughout the prey's life cycle, amplifying the complexity of their ecological dynamics. In their experiment involving song sparrows, Zanette et al. [25] observed a 40% reduction in the breeding rate due to the induced fear of predation risk. In another independent study, Elliott et al.[26] reported that Drosophila melanogaster exhibits anti-predator behaviors upon exposure to Mantid odor, displaying reduced activity across both breeding and non-breeding seasons. Meanwhile, Wang et al. [27] presented a mathematical model for prey-predator dynamics, incorporating the fear risk. They disclosed that heightened fear levels might stabilize the ecosystem by preventing the occurrence of periodic oscillations in the population. For more information about predator-prey models with fear effects, see [28,29,30] and references cited therein. The fear effect is not limited to only common symbiosis relationships like mutualism, competition and predator/prey interactions. It has also been observed in commensalism and amensalism types of symbiosis. For example, Xi et al. [2] studied the number of cocoons formed by grassland caterpillars in the presence or absence of grasshoppers and obtained Figure 1. From Figure 1, it can be clearly observed that the fear of grasshoppers by caterpillars can lead to a decline in species numbers. Ogada [31] have highlighted that the presence of large herbivorous mammals in African savanna habitats has resulted in a reduction of the canopy area of trees, as well as a decrease in grass cover by 8% and forb cover by 33%. However, the canopy serves as a crucial habitat and food source for birds in this particular East African savanna ecosystem. Thus, the sizable herbivore has exerted a significant influence on avian behavior and growth, particularly in relation to foraging habits, rates of maturation, chances of survival, reproductive capacity and dietary resources. As a result, the overall diversity of bird species has experienced a decline of approximately 30%. This decline may be interpreted as the fear induced by large herbivores in the East African savanna ecosystem, resulting in reduced avian birth rates. Consequently, developing models that can effectively explain this phenomenon is imperative. However, no scholarly investigations have taken into account the impact of the fear effect on the amensalism system.
Recently, discrete-time models have attracted considerable attention due to their rich and complex dynamic behaviors and have been explored extensively [18,30,32]. However, there needs to be more emphasis on investigating discrete-time amensalism systems. In 2022, Zhou et al. [21] investigated the following discrete amensalism systems:
{xn+1=xnexp(α−βxn−cyn),yn+1=ynexp(γ−δyn). | (1.2) |
They studied the stability and attractivity of the equilibrium point of system (1.2) and obtained the following results:
Theorem B (1) If 0<α≤2, 0<c<δαγ, 0<γ<2 or α>2, δ(α−2)γ<c<δαγ, 0<γ<2, then the positive equilibrium point E∗3(x∗3,γδ) is local stability.
(2) If 0<γ<2, 0<α−cγδ<ln2+1, then the positive equilibrium point E∗3(x∗3,γδ) is globally attractive.
Obviously, system (1.2) is the discrete form of model (1.1). In system (1.1), according to Theorem A (2), when a positive equilibrium point exists, it is global stability. This also implied that no matter what the initial conditions are, the solution will eventually converge to the positive equilibrium point. While for discrete system (1.2), Theorem B shows that the conclusion of the stability of the positive equilibrium point can only be obtained under stricter conditions than Theorem A. Additionally, model (1.2) undergoes flip bifurcation at the boundary equilibrium points E1, E2 and the positive equilibrium point E∗3 [21]. This is a dynamic phenomenon that does not exist in the continuous system (1.1), Hence, the investigation of discrete models is highly valuable due to their propensity for exhibiting intricate dynamic phenomena that surpass those observed in continuous models.
Subsequently, Zhou et al.[32] investigated a discrete amensalism system with the Beddington-DeAngelies functional response and Allee effect, and they found that the Allee effect causes the system to take longer trajectories to reach its stable steady-state solution. So far, to the best of our knowledge, there has been no research conducted on the discrete amensalism system with the fear effect by researchers.
Motivated by prior research, we assume that the birth rate of the first species will be reduced by the fear of the second species. We multiply the birth rate of the first species by a decreasing function G(k,y) related to the size of the second species. Based on observations in the biological world, the fear function G(k,y) proposed by Wang et al. [27] must satisfy the following requirements:
G(0,y)=1,G(k,0)=1,limy→∞G(k,y)=0,limk→∞G(k,y)=0,∂G(k,y)∂k<0,∂G(k,y)∂y<0, |
where k represents the level of fear and y is the species density. Note that G(k,y)=1/(1+ky) meets the above requirements and 1/(1+ky) is now widely applied in [33,34,35].
Accordingly, on the basis of system (1.1), considering the above assumption, we introduce the fear effect to the first species, resulting in the formulation of the following amensalism model:
{dxdt=x(e11+k1y−e2−b1x−c1y),dydt=y(a2−c2y), | (1.3) |
where e1 and e2 represent the birth and death rates of the first species and 1/(1+ky) represents the fear effect function. It is to be noted that without the fear effect (k=0), the above model (1.3) will degenerate to the amensalism model (1.1). For the convenience of analysis, with the transformation x=ˉx/b1, y=ˉy/c2, t=ˉt, after dropping the bars, system (1.3) becomes
{dxdt=x(e11+ky−e2−x−cy),dydt=y(a2−y), | (1.4) |
where k=k1/c2 and c=c1/c2. The numerical simulations in [36] show that when the step size gets bigger in Euler's scheme, both period-doubling and Neimark-Sacker bifurcation happen. This means that the numerical approach for discretization is not accurate. In order to address this shortcoming, the method of piecewise constant arguments was established. Based on the findings in [37,38], the method of piecewise constant arguments is indeed a preferable solution for the discretization of continuous models. Hence, using the method of piecewise constant arguments proposed by Jiang and Rogers [39], similar to the modeling method in [32], we can obtain the following discrete-time model:
{xn+1=xnexp(e11+kyn−e2−xn−cyn),yn+1=ynexp(a2−yn). | (1.5) |
Obviously, the initial conditions (x0,y0) for (1.5) are in R2+. In a discrete amensalism ecological model with the fear effect, what kind of dynamical behavior does it exhibit? Does the stability of the equilibrium point change compared to a continuous system (1.1)? Does a new bifurcation happen compared to a discrete system (1.2)? What effects do fear effects have on the system? We will conduct a detailed investigation to answer these questions.
The layout of the paper is as follows: In Section 2, we analyze the relevant properties of the fixed points. Section 3 investigates bifurcation phenomena and considers the chaos control method. Then, in Section 4, we use numerical simulations to validate the feasibility of the obtained results. Finally, a brief discussion is provided at the end of the paper.
Note that fixed points are determined by the following system of equations:
{x=xexp(e11+ky−e2−x−cy),y=yexp(a2−y). | (2.1) |
Obviously, system (1.5) always has four fixed points: E0(0,0), E1(0,a2), E2(e1−e2,0) and E∗((e1−e2−ke2a2−ka22c−a2c)/(1+ka2),a2). The fixed points E2 and E∗ are non-negative if e2<e1 and e1>(a2k+1)(a2c+e2) are satisfied, respectively. There are two fixed points in region A of Figure 2(a): E0 and E1. There are three fixed points in region B: E0, E1 and E2. There are four fixed points in region C: E0, E1, E2 and E∗.
The Jacobian matrix of system (2.1) is calculated as follows:
J(x,y)=((1−x)M−x(e1k(1+ky)2+c)M0(1−y)N), | (2.2) |
where M=exp(e1/(1+ky)−e2−x−cy) and N=exp(a2−y).
Next, we analyze the local stability of the equilibria according to the above Jacobian matrix.
Theorem 2.1. E0(0,0) is
1) a source if 0<e2<e1;
2) a saddle if 0<e1<e2;
3) non-hyperbolic if e1=e2.
Proof. For the equilibrium E0(0,0), the Jacobian matrix is
J(E0)=(exp(e1−e2)00exp(a2)), |
with eigenvalues λ1=exp(e1−e2)>0 and λ2=exp(a2)>1. Thus, E0(0,0) is a source if 0<e2<e1, a saddle if 0<e1<e2 and non-hyperbolic if e1=e2. The proof is completed.
Theorem 2.2. The local stability of E1(0,a2) is described below:
1) a sink if 0<e1<(a2k+1)(a2c+e2) and 0<a2<2;
2) a source if e1>(a2k+1)(a2c+e2) and a2>2;
3) non-hyperbolic if e1=(a2k+1)(a2c+e2) or a2=2;
4) a saddle if one of the conditions holds:
(a) 0<e1<(a2k+1)(a2c+e2) and a2>2;
(b) e1>(a2k+1)(a2c+e2) and 0<a2<2.
Proof. The Jacobian matrix of system (2.1) at E1(0,a2) is given by
J(E1)=(exp(e11+ka2−e2−ca2)001−a2). |
Consequently, two eigenvalues are λ1=exp(e1/(1+ka2)−e2−ca2)>0 and λ2=1−a2<1. Hence, if 0<e1<(a2k+1)(a2c+e2) and 0<a2<2, then we have 0<λ1<1 and |λ2|<1. Therefore, E1(0,a2) is a sink. The proofs of (2–4) are similar to (1) and are omitted here. The topological classification of E1 is presented in Figure 2(b).
Theorem 2.3. E2(e1−e2,0) is
1) a source if e1>e2+2;
2) a saddle if 0<e1<2+e2;
3) non-hyperbolic if e1−e2=2.
Proof. For the equilibrium E2(e1−e2,0), the Jacobian matrix is
J(E2)=(1−e1+e2e1−e2(−e1k−c)0exp(a2)). |
Obviously, the two eigenvalues are λ1=1−e1+e2=1−(e1−e2)<1 and λ2=exp(a2)>1. If e1−e2>2, i.e., λ1<−1, then |λ1|>1 and |λ2|>1, implying that E2(e1−e2,0) is a source. Therefore, (1) holds. The proof processes of (2) and (3) are similar, we omit the details here. Thus, E2 is always unstable. This completes the proof.
For E∗(x∗,a2), the Jacobian matrix is as follow:
J(E∗)=(1−x∗x∗(e1k(1+ka2)2+c)01−a2), |
where λ1=1−x∗=1−(e1−(1+a2k)(e2+ca2))/(1+ka2)<1 and λ2=1−a2<1. Note that when 0<c<(e1−e2(1+ka2))/(1+ka2)a2≜˜c and 0<e2<e1/(1+ka2)−2≜e∗∗2. In addition, we can verify c1<˜c. Then,
1−e1−(1+a2k)(e2+ca2)1+ka2{<−1if 0<c<c1,=−1if c=c1,∈(−1,1)if c1<c<˜c, |
where c1=(e1−(2+e2)(1+ka2))/(1+ka2)a2. Moreover, if e∗∗2<e2<e1/(1+ka2) and 0<c<˜c, then 1−x∗∈(−1,1). Therefore, the following result can be obtained immediately.
Theorem 2.4. Suppose that 0<c<˜c. Then, the topological classifications of the unique positive fixed point E∗(x∗,a2) is given by Table 1.
Conditions | Case | ||
0<a2<2 | sink | ||
e∗∗2<e2<e11+ka2 | 0<c<˜c | a2>2 | saddle |
a2=2 | non-hyperbolic | ||
0<a2<2 | saddle | ||
0<c<c1 | a2>2 | source | |
a2=2 | non-hyperbolic | ||
0<a2<2 | non-hyperbolic | ||
0<e2<e∗∗2 | c=c1 | a2>2 | non-hyperbolic |
a2=2 | non-hyperbolic | ||
0<a2<2 | sink | ||
c1<c<˜c | a2>2 | saddle | |
a2=2 | non-hyperbolic |
In this section, we consider the global asymptotic stability of the equilibrium point with the help of Lemmas 3–5 from [30] and Lemmas 2.1 and 2.2 from [40].
Theorem 2.5. Suppose that 0<a2<2, 0<k∗∗<k<k∗ holds and E∗(x∗,a2) is globally attractive, i.e.,
limn→∞x(n)=x∗,limn→∞y(n)=a2, | (2.3) |
where k∗=e1−e2−ca2(e2+ca2)a2 and k∗∗=e1−ln2−1−e2−ca2(ln2+1+e2+ca2)a2.
Proof. According to Lemma 4 in [30], if 0<a2<2 holds, then
limn→∞y(n)=a2. |
For sufficiently small ε, there exists an integer N1, such that for all n>N1,
y(n)>a2−ε. | (2.4) |
From Lemma 2.1 and 2.2 in [40], we obtain:
lim infn→+∞y(n)≥min{a2exp(a2−exp(a2−1)),a2}:=H1. | (2.5) |
Now, we consider the following system:
x1(n+1)=x1(n)exp(e11+ka2−e2−x1(n)−ca2). | (2.6) |
Apparently, x∗1=x∗ is the positive equilibrium point of the system. Suppose {x1(n)} is any positive solution of system (1.5). From Lemma 3 in [30], if 0<e1−(1+ka2)(e2+ca2)<2, we know that limn→∞x1(n)=x∗. Thus, we only need to prove that
limn→∞(x(n)−x1(n))=0. |
Assume that
x(n)=x1(n)exp[k1(n)]. |
Therefore, our main goal is to prove limn→∞k1(n)=0.
The first equation of (1.5) is equivalent to
k1(n+1)=lnx(n)+e11+ky(n)−e2−x(n)−cy(n)−lnx1(n+1)=k1(n)(1−x1(n)exp[θ1(n)k1(n)])−(c+ke1(1+a2k)(1+ky(n)))(y(n)−a2). | (2.7) |
Then, x1(n)exp[θ1(n)k1(n)] is between x1(n) and x(n), and here θ1(n)∈[0,1].
It follows from (2.5) that
ke1(1+a2k)(1+ky(n))≤ke1(1+a2k)(1+kH1):=W1. |
Obviously,
x(n)exp[e11+(a2+ε)k−e2−c(a2+ε)−x1(n)]≤x(n+1)≤x(n)exp[e11+a2k−e2−ca2−x1(n)], |
holds. Thus, according to Lemma 2.1 and 2.2 in [40], it immediately follows that:
lim supn→+∞x(n)≤exp(e11+a2k−e2−ca2−1):=Q1,lim infn→+∞x(n)≥min{(e11+(a2+ε)k−e2−c(a2+ε))exp[−e2+e11+(a2+ε)k−c(a2+ε)−Q1],e11+(a2+ε)k−e2−c(a2+ε)}:=M1. |
Setting ε→0 in the above inequalities leads to
lim supn→+∞x1(n)≤exp(e11+a2k−e2−ca2−1)=Q1,lim infn→+∞x1(n)≥min{(e11+a2k−e2−ca2)exp[e11+a2k−e2−ca2−Q1],e11+a2k−e2−ca2}≥M1. |
Hence, taking an integer N2>N1, for ξ>0 small enough, when n≥N2, the following inequality holds:
M1−ξ≤x(n),x1(n)≤Q1+ξ,n≥N2. | (2.8) |
Assume
λ1=max{|1−M1|,|1−Q1|}. |
Now, let
λε1=max{|1−(M1−ξ)|,|1−(Q1+ξ)|}. | (2.9) |
Therefore,
|k1(n+1)|≤max{|1−(M1−ξ)|,|1−(Q1+ξ)|}|k1(n)|+(c+W1)ε=λε1+(c+W1)ε,n≥N2. |
The following equation can be obtained:
|k1(n)|≤λn−N2ε1|k1(N2)|+1−λn−N2ε11−λε1(c+W1)ε,n≥N2. | (2.10) |
Considering λε1<1, then limn→+∞k1(n)=0, i.e., limn→+∞[x(n)−x1(n)]=0 is established if λ1<1. Since
1−Q1<1−M1<1, |
λ1<1 is equivalent to 1−Q1>−1, i.e.,
e11+a2k−e2−ca2<1+ln2. |
Hence, Theorem 3.1 is proved.
Now, we discuss the global attractivity of E1 with the help of Lemma 6 and 7 from [30].
Theorem 2.6. Assuming that k>k∗, 0<a2<2 holds, and E1(0,a2) is globally attractive, i.e.,
limn→∞x(n)=0,limn→∞y(n)=a2. |
Proof. The proof of limn→∞y(n)=a2 follows the same process as the one in Theorem 2.5 above, so it is omitted here. Next, we prove that limn→∞x(n)=0.
Consider the auxiliary equation:
x2(n+1)=x2(n)exp(e11+ka2−e2−x2(n)−cy(n)), |
combined with the second equation of system (1.5), for all i∈N, we have
lnx2(i+1)x2(i)=e11+ka2−e2−x2(i)−cy(i)lny(i+1)y(i)=a2−y(i). | (2.11) |
According to the Theorem 2.6, we have e11+ka2−e2a2<c, i.e.,
e11+ka2−e2−ca2<0. | (2.12) |
Therefore, the following inequality holds:
e11+ka2−e2a2<sh<c, |
where s and h are positive numbers. Then,
hc−s>0 | (2.13) |
and there exist δ>0 such that
h(e11+ka2−e2)−sa2<−δ<0. | (2.14) |
From (2.11)–(2.14), we have
hlnx2(i+1)x2(i)−slny(i+1)y(i)≤[h(e11+ka2−e2)−sa2]−[hx2(i)+(hc−s)y(i)]≤h(e11+ka2−e2)−sa2<−δ<0. |
Summing up the above formulas, we can get:
n∑i=1(hlnx2(i+1)x2(i)−slny(i+1)y(i))=hlnx2(n)x2(0)−slny(n)y(0)<−δn. |
Then,
x2(n)<x2(0)(y(n)y(0))shexp(−δhn). | (2.15) |
According to Lemma 2.1 in [40] and y(n+1)≤y(n)exp(a2−y(n)), it holds that:
lim supn→+∞y(n)≤exp(a2−1). |
Therefore, taking an integer N2>N1, for ε>0, when n≥N2, y(n)≤exp(a2−1)+ε:=M is established. So, the inequality (2.15) is equivalent to
x2(n)<x2(0)(My(0))shexp(−δhn),∀n≥N2. |
Thus,
limn→∞x2(n)=0. | (2.16) |
Therefore, if we want to prove limn→∞x(n)=0, we only need to prove
limn→∞(x(n)−x2(n))=0, |
Assume that
x(n)=x2(n)exp[k2(n)]. |
Therefore, we can convert what we need to prove into the proof limn→∞k2(n)=0.
By a similar process of proof as in Theorem 2.5, we have that when
e11+ka2−e2−ca2<ln2+1, | (2.17) |
limn→∞k2(n)=0 holds. Combining condition (2.11) and condition (2.16), we have that limn→∞x(n)=0 holds when e1/(1+ka2)−e2−ca2<0. Hence, Theorem 2.6 is proved.
Based on the above proofs, it becomes evident that the system (1.5) exhibits several bifurcation types at its fixed points. In the subsequent discussion, we will examine these potential bifurcations by employing the center manifold [41] and bifurcation theory [42,43].
Theorem 2.2 states that if a2=2 and e1≠(1+a2k)(ca2+e2) holds, then one of the eigenvalues of E1(0,a2) is -1 and another one is neither 1 nor –1. These requirements suggest that all parameters belong to the set TA:
TA:={(e1,e2,k,c,a2):a2=2,e1≠(1+a2k)(ca2+e2),e1,e2,c>0,k≥0}. |
At this point, the central manifold of system (1.5) at E1(0,a2) is x=0. Therefore, the restricted system of (1.5) to it is:
yn+1=f(yn)=ynexp(a2−yn). |
f′(a2)=−1 is obvious, so it can be seen that a flip bifurcation occurs at E1(0,a2). Figure 3(a) depicts the bifurcation diagram about E1.
Based on conclusion (4) of Theorem 2.3, we obtain that all parameters belong to the set TB:
TB:={(e1,e2,k,c,a2):e1=e2+2,a2,e2,c>0,k≥0}. |
We can easily know that the central manifold of system (1.5) at E2(e1−e2,0) is y=0. So the restricted system of (1.5) to it is:
xn+1=g(xn)=xnexp(e1−e2−xn). |
Thus, we have g′(e1−e2)=−1. Figure 3(b) depicts the bifurcation diagram about E2. Moreover, as illustrated in Figure 3, the maximum Lyapunov exponents (MLE) for(a), (b) are depicted in (c) and (d), respectively.
According to Theorem 2.4, if e∗1=(1+ka2)(2+e2+a2c), then at E∗(x∗,a2), one eigenvalue is -1 and another one is neither 1 nor -1, These conditions imply that all parameters belong to the set Fc:
Fc:={(e∗1,e2,k,c,a2):e∗1=(1+ka2)(2+e2+a2c),a2≠2,e2,c,a2>0,k≥0}. |
If the condition of (e∗1,e2,k,c,a2)∈Fc is satisfied, flip bifurcation may occur. Thus, suppose δ is a bifurcation parameter satisfying ‖δ‖≪1. Then, system (1.5) can be expressed as follows:
(xy)→(xexp(e∗1+δ1+ky−e2−x−cy)yexp(a2−y)). | (3.1) |
First, we transform E∗ to the origin by using the transformation u=x−x∗,v=y−a2 and then utilize Taylor expansion, turning the mapping (3.1) into
(uv)→(−1a01001−a2 )(uv)+(f3(u,v,δ)g3(u,v,δ)), | (3.2) |
where
f3(u,v,δ)=z101uv+s110uδ+z011v2+s220vδ+z022u2δ+s330u3+z033uv2+s440uvδ+s550v3+z055v2δ+s660vδ2+O((|u|+|v|+|δ|)4),g3(u,v,δ)=(a2−2)v22−(a3−3)v36+O((|u|+|v|+|δ|)4) |
and
a010=−2z101,z101=2a2ck+e2k+c+2ka2k+1,s110=−1a2k+1,z011=z2101+2k2(a2c+e2+2)a2k+1,s220=−z011a2k+1−2k(a2k+1)2,s330=16,z022=12(a2k+1),z033=z0112,s440=z011a2k+1+k(a2k+1)2,s550=−z011z1013−z2101(4a2ck2+6e2k2+11k23(1+a2k)−k2(−2e2k−2e2c+11e2k−3c+6k)3(1+a2k)3,z055=z0112(a2k+1)+2ck(1+a2k)2−2k(a2c+e2+3)(1+a2k)3,s660=−k(a2k+1)2. |
Next, we use the invertible transformation:
(uv)=(a010a01002−a2)(XY). |
Equation (3.2) can be rewritten as
(XY)→(−1001−a2)(XY)+(f4(u,v,δ)g4(u,v,δ)), | (3.3) |
where
u=a010(X+Y),v=(2−a2)Y,f4(u,v,δ)=b101uv+a110uδ+b011v2+a220vδ+a330u3+b033uv2+a440uvδ+a550v3+b550v2δ+a660vδ2+b066u2δ+O((|u|+|v|+|δ|)4),g4(u,v,δ)=−12v2−a3−36(a2−2)v3+O((|u|+|v|+|δ|)4) |
and
b110=z101a010,a110=s110a010,b011=z011a010+12,a220=s220a010,a330=s330a010,b033=z033a010,a440=s440a010,a550=s550a010−a2−36(a2−2),b550=z055a010,a660=s660a010,b066=z022a010 |
According to the central manifold theorem, there exists a central limit Wc1(0,0,0), which can be expressed as
Wc1(0,0,0)={(X,Y,δ):Y=h(X,δ),h(0,0)=0,Dh(0,0)=0}, |
where
h(X,δ)=t1X2+t2Xδ+t3δ2+O((|X|+|δ|)3. | (3.4) |
Based on (3.3), Y=h(X,δ) must satisfy:
h(−X+f4(X,h(X,δ),δ),δ)=(1−a2)h(X,δ)+g4(X,h(X,δ),δ). |
By calculation, we obtain t1=t2=t3=0. Thus, the map (3.3) can be written as :
G∗:X→−X+c110δ+c011Xδ+c220δ2+c330X3+c033Xδ2+c440δ3+O((|X|+|δ|)3), |
where
c110=2a010(ka2+1)δ,c011=−1ka2+1,c220=1(ka2+1)2a010,c330=a20106,c033=−12(ka2+1)2,c440=13(ka2+1)3. |
Then, we have:
ϖ1=(G∗Xδ+12G∗δG∗XX)|(X,δ)=(0,0)=−1ka2+1≠0,ϖ2=(16G∗XXX+(12G∗XX)2)|(X,δ)=(0,0)=a20106≠0. |
Consequently, the following conclusions can be drawn.
Theorem 3.1. If (e∗1,e2,k,c,a2)∈Fc, then the system (1.5) undergoes flip bifurcation at E∗(x∗,a2). (see Figure 7)
According to Theorem 2.1, if e1=e2, then at E0(0,0), one eigenvalue is 1 and another one is neither 1 nor -1. These conditions mean that all of the parameters are in the set Fd:
Fd:={(e1,e2,k,c,a2):e1=e2,e1,e2,c,a2>0,k≥0}. |
Taking e1 as the bifurcation parameter and setting η as the disturbance parameter, the map is as follows:
(xy)→(xexp(e∗1+η1+ky−e2−x−cy)yexp(a2−y)). | (3.5) |
Next, a Taylor expansion is performed at the origin. Then, (3.5) becomes
(xy)→(100λ2)(xy)+(f(x,y,η)g(x,y,η)), | (3.6) |
where
f(x,y,η)=−x2+xη−(e2k+c)yx+x32−x2η+(e2k+e1)yx2+xη22(e22k2+2ce2k+2e2k2+c2)y2x2−(e2k+c+k)yxη+O((|x|+|y|+|η|)4),g(x,y,η)=−λ2y2+λ22y3+O((|x|+|y|+|η|)4), |
Setting the approximate central mainfold Wc2(0,0,0) of (3.6) as
Wc2(0,0,0)={(x,y,η):y=h1x2+h2xη+h3η2+O((|x|+|η|)3)}, |
for x and η sufficiently small. By a simple coefficient comparison, we can obtain h1=h2=h3=0. Therefore, the map on the center manifold is:
K:x→x−x2+xη+12x3−x2η+12η2x+O((|x|+|η|)4). |
By simple compulation, we have:
{L1Δ=(∂2K∂x2)(0,0)=−2≠0,L2Δ=(∂2K∂x∂η)(0,0)=1≠0. |
Hence, we draw the following conclusion about transcritical bifurcation.
Theorem 3.2. If (e1,e2,k,c,a2)∈Fd, then system (1.5) experiences transcritical bifurcation at E0(0,0).
Considering that the flip bifurcation may lead to chaos, in this section, we focus on controlling the generation of chaos. Combined with the method of hybrid control, the related control system (1.5) is expressed as follows:
{xn+1=ρxnexp(e11+ky−e2−x−cyn)+(1−ρ)xn,yn+1=ρynexp(a2−yn)+(1−ρ)yn, | (3.7) |
where ρ∈(0,1). The Jacobian matrix of E∗(x∗,a2) in the control system (3.7) is as follows:
J(E∗)=(1−ρ(e1−(1+a2k)(ca2+e2))1+ka2−ρx∗(e1k(1+ka2)2)01−ρa2). | (3.8) |
Its eigenvalues are λ1=1−ρ(e1−(1+a2k)(ca2+e2))/(1+ka2)<1 and λ2=1−ρa2<1, respectively. Hence, we have the following outcome:
Theorem 3.3. The controlled system (1.5) is locally asymptotically stable at E∗(x∗,a2) when
0<ρ<min{2a2,2(1+ka2)e1−(1+a2k)(ca2+e2),1}. |
In this section, we will illustrate the accuracy of relevant conclusions and explore the influence of the fear effect on the system through numerical simulations.
Example 1. (the effect of fear effect on E1)
1) Given the values (e1,a2,e2,c,k)=(1,1.7,0.4,0.3,0.1), it can be observed that (a2k+1)(a2c+e2)=1.06, which is larger than e1. Additionally, it is known that 0<a2<2. Consider the three sets of initial values: (0.3,0.8), (0.6,0.2) and (1.2,0.7). As depicted in Figure 5(a), it can be observed that E1 exhibits local asymptotic stability. The correctness of the result of Theorem 3 is affirmed. Furthermore, to investigate the impact of the fear effect on the local stability of E1, we varied the parameter k from 0.1 to 5. The findings are depicted in Figure 5(b). It can be observed that the presence of fear induces an increase in the local stability of E1, thereby expediting the process of extinction for the first species.
2) Given the values (e1,e2,c,k)=(4,0.5,1,2) and the condition a2>2, we observe that e1=4<(a2k+1)(a2c+e2). satisfying condition (3) of Theorem 2.2 and indicating that E1 is unstable at this particular moment. Next, we select a2 as the bifurcation parameter to investigate the impact of a2 on the system. As depicted in Figure 6, when the value of a2 grows, the first species will undergo extinction, whereas the second species will experience a state of chaos. Furthermore, the aforementioned conditions fail to meet the requirements stated in Theorem 2.6, namely the conditions for achieving global stability of E1. Therefore, this example also provides additional evidence about the plausibility of the conditions outlined in Theorem 2.6.
Example 2. (the fear effects on the first species)
Case 1: Assuming that the second species is global stability, i.e., 0<a2<2
When the value of k is set to 0.3, Figure 7(a) illustrates periodic-2, 4 and 8 orbits. To provide a more intuitive observation of periodic orbits and chaotic sets, we present the phase diagrams corresponding to each stage of Figure 7(a) for six different values of e1: 2,3.25,3.6,4.1,4.2,4.3. (as shown in Figure 8). The first species exhibits stability when e1 falls between 0 and 3.25, becoming unstable when e1 exceeds 3.25. It is observed that with an increase in the birth rate, the initial species exhibits instability or even chaotic behavior. In this study, our goal is to investigate the impact of the fear effect on the first species under conditions of chaos. We set the parameter e1 to a value of 4.3 and utilize the bifurcation parameter k. The stability of E∗(x∗,a2) is observed to be dependent on the value of k. Specifically, E∗(x∗,a2) is unstable for 0<k<0.65 and stable for k>0.65. The analysis of Figure 7(b) reveals that the fear effect significantly contributes to preserving stability in the first species, especially when the second species achieves global stability.
Case 2: Assuming that the second species is chaos, i.e., a2>2.
When a2>2, the second species exhibits chaotic behavior, leading to the following scenarios. In this chaotic state, it is observed that the fear effect no longer has the effect of stabilizing the first species. Moreover, with the increase of the parameter k, the density of the first species decreases, eventually leading to its extinction when k>k∗.
Figure 10 depicts the effect of the parameters e1 and k, as well as the parameters a2 and k, on the bifurcation.
Furthermore, when the parameter k is set to zero, a comparison between system (1.1) and system (1.5) reveals that the variable a1 in system (1.1) corresponds to the difference between e1 and e2 in system (1.5). Additionally, both b1 and c2 in system (1.1) assume a value of 1 in system (1.5). Consequently, when the parameter values of system (1.5) are set to (a2,c,e2,e1)=(3,0.2,0.3,2.3), it corresponds to the parameter values of system (1.1) being (a2,c1,a1,c2,b1)=(3,0.2,2,1,1). At this time, it is observed that a1/c1>a2/c2. According to the findings of Theorem A, it can be deduced that the positive equilibrium point in system (1.1) is globally stable. However, the positive equilibrium point in system (1.5) is unstable, and there exists a potential for flip bifurcation. Consequently, the discrete model exhibits more intricate dynamical behavior in comparison to the continuous model.
Example 3. (the effect of fear effect on E∗)
We examine the stability of the positive equilibrium point E∗(x∗,a2). The parameters are assigned values of (a2,e2,c,e1)=(0.5,1,0.2,3), and the initial values (x0,y0) are set to (0.4,0.3), (0.5,0.8) and (1.1,0.6) accordingly. Based on the conditions stated in Theorem 2.5, it can be concluded that when k∈(0.07,3.4545), the positive equilibrium point E∗(x∗,a2) is globally stable. To verify the feasibility of Theorem 2.5, we select k=2, as depicted in Figure 12(a). It can be observed that at this value of k, the equilibrium point E∗ remains globally stable. Furthermore, in order to investigate the effect of the fear effect on the global stability of E∗, we chose k=4. The results show that the fear effect affects the global stability of the positive equilibrium point and reduces the density of the first species. This reduction in density eventually leads to the extinction of the first species.
The values assigned to (a2,e2,c,e1) are (1,0.5,0.3,4). For this set of parameters, the values of k∗∗ and k∗ are 0.60 and 4, respectively. We will use k as the bifurcation parameter to investigate the influence of k on the stability of E∗. At this time, the second species is stable. Combined with Figure 13(a), we can see that when k∗∗<k<k∗, E∗ is stable; when k>k∗, the first species becomes extinct. Therefore, we can conclude that an appropriate value of k promotes the convergence of the first species from a chaotic state to stability.
Additionally, examining the two trajectories shown in Figure 13(b) with k=0 and k=1 makes it clear that, as k increases, the solution approaches a stable state within shorter trajectories. This observation underscores the notion that an appropriate level of fear might confer advantages to species, enabling them to acclimate more effectively to changes in the surrounding environment.
Example 4. (Effect of fear effect on the stability of the system)
We proceed to investigate the influence of the fear effect on the system through numerical simulations. The chosen parameter values for this analysis are (e1,a2,e2,c)=(2,1,0.3,0.2). The density distribution of the first species, denoted as x, and the second species, denoted as y, for varying values of k is illustrated in Figure 14. In this analysis, we examine two scenarios: one where the system exhibits a fear effect (k≠0) and another where there is no fear effect (k=0). Furthermore, we categorize the presence of the fear effect into several levels, specifically k=0.3,k=0.5,k=1,k=3,k=5. It was observed that an increase in the fear effect parameter k reduced the density of the first species and had no effect on the second species. However, it enhances the local stability of the first species and diminishes the frequency of oscillations. As the intensity of fear increases, the local stability of the first species will diminish, resulting in the prolonged period required to attain equilibrium. Moreover, a higher value of k leads to the extinction of the first species.
Example 5. The bifurcation parameter for the chaotic region is denoted as k, while the remaining parameter values are established as follows: the values of the variables (e1,a2,e2,c) are as follows: e1=4.3, a2=1, e2=0.3, c=0.2, the given system, denoted as Eq (3.7), can be expressed as:
{xn+1=ρxnexp(41+kyn−0.3xn−0.2yn)+(1−ρ)xn,yn+1=ρynexp(1−yn)+(1−ρ)yn. | (4.1) |
Based on the observation of Figure 7, it is evident that for values of k within the range of 0<k<0.5, the first species undergoes a chaotic state. Therefore, we select certain values of k such as k=0.20,0.25,0.30,0.35,0.40,0.45. By utilizing Theorem 3.3, in conjunction with Table 2, we can determine the range of values for ρ when 0.2≤k≤0.5.
Values of bifurcation parameter k from the chaotic region | Stability interval ρ |
0.20 | 0<ρ<0.648648648648 |
0.25 | 0<ρ<0.680272108843 |
0.30 | 0<ρ<0.712328767123 |
0.35 | 0<ρ<0.744827586206 |
0.40 | 0<ρ<0.777777777777 |
0.45 | 0<ρ<0.811188811188 |
Figure 15(a), (b) show that chaos control is effective.
Through numerical simulations, we find that when the fear effect k changes, the dynamic behaviors of the system will also change accordingly. We summarize the impact of the fear effect k on the system in Table 3:
Conditions | Case | |
k>k∗ | The first species became extinct. | |
0<a2<2 | As k increases, the density of the first species decreases. | |
k∗∗<k<k∗ | When the first species is in chaos, k has a stabilizing effect. | |
k>k∗ | The first species became extinct. | |
a2>2 | 0<k<k∗ | Decrease in the density of the first species. |
Additionally, through numerical simulations, we have observed some intuitively understandable phenomena:
1) Compared with the model (1.1) studied by Sun [9], considering the discrete amensalism model (1.5) with fear effect on the first species, the stability of the positive equilibrium point in model (1.5) changes, and the dynamic behavior will become more complicated. For example, transcritical bifurcation and flip bifurcation appear.
2) Compared to system (1.2), the trajectory of system (1.5) reaches its stable steady-state solution more rapidly as the fear effect increases. However, the first species will go extinct more quickly with an increasing fear effect.
3) The dynamic behavior of system (1.5) is different from that of system (1.2). For example, there is a flip bifurcation at the boundary equilibrium point and the positive equilibrium point, but there is also a transcritical bifurcation at E0(0,0), which is not seen in system (1.2). The presence of a transcritical bifurcation indicates that system (1.5) has some plasticity and adaptability to change from one stable state to another in the face of external changes or stresses. This adaptability helps species adapt more quickly to changes in their environment.
4) Zhou [21] did not study the global attractivity of the boundary equilibrium point of system (1.2). However, we notice that the boundary equilibria of system (1.5) is also globally attractive and its attractivity depends on the fear effect k, which means that the second species is stable and the extinction of the first species depends on the fear effect.
5) By observing Figure 1, Xi et al. [2] can only conclude that the fear effect will reduce the density of species. However, they have not conducted in-depth research on how the fear effect will be affected if it gradually increases. Through numerical simulation, it is not difficult to find if the fear effect is large enough to cause the extinction of the first species.
Finally, we must point out that an appropriate model usually takes functional responses into account, and the recent paper [18] explored the dynamical behavior of amensalism system with Beddington- DeAngelies functional response. However, no scholars have yet explored the influence of the fear effect on the dynamical behavior of amensalism system with Beddington-DeAngelies functional response, which will be further investigated in our subsequent work.
The authors declare that they have not used artificial intelligence tools in the creation of this article.
The authors would like to thank four anonymous reviewers for their valuable comments, which greatly improved the final expression of the paper.
The authors declare that there is no conflict of interest.
Lemma 3 Let f(u)=uexp(α−βu), where α and β are positive constants, then f(u) is nondecreasing for u∈(0,1β].
Lemma 4 Assume that sequence {un} satisfy un+1=unexp(α−βun), n∈N, where α and β are positive constants and u(0)>0. Then
(i) if α<2, then limn→+∞u(n)=αβ.
(ii) if α≤1, then u(n)≤1β,n=2,3,....
Lemma 5 Suppose that function f,g:Z+×[0,+∞)→[0,+∞) satisfy f(n,x)≤g(n,x)(f(n,x)≥g(n,x)) for n∈Z+ and x∈[0,+∞) and g(n,x) is nondecreasing with respect to x. If {x(n)} and {u(n)} are the solutions of the following difference equations
x(n+1)=f(n,x(n)), u(n+1)=g(n,u(n)), |
respectively, and x(0)≤u(0)(x(0)≥u(0)), then
x(n)≤u(n)(x(n)≥u(n)),∀n≥0. |
Lemma 2.1 Assume that {x(k)} satisfies x(k)>0 and
x(k+1)≤x(k)exp[a(k)−b(k)x(k)],k∈N, |
where a(k) and b(k) are non-negative sequences bounded above and below by positive constants. Then
lim supk→+∞x(k)≤exp(au−1)bl, |
where au=supn∈N{a(n)},bl=infn∈N{b(n)}.
Lemma 2.2 Assume that {x(k)} satisfies
x(k+1)≥x(k)exp[a(k)−b(k)x(k)],k≥N0, |
lim supk→+∞x(n)≤x∗, and x(N0)>0, where a(k) and b(k) are non-negative sequences bounded above and below by positive constants and N0∈N. Then
lim infn→+∞x(n)≥min{albuexp(al−bux∗),albu}, |
where al=infn∈N{a(n)},bu=supn∈N{b(n)}.
[1] |
J. V. Corte-Real, H. M. Baldauf, J. Abrantes, P. J. Esteves, Evolution of the guanylate binding protein (GBP) genes: Emergence of GBP7 genes in primates and further acquisition of a unique GBP3 gene in simians, Mol. Immunol., 132 (2021), 79–81. https://doi.org/10.1016/j.molimm.2021.01.025 doi: 10.1016/j.molimm.2021.01.025
![]() |
[2] |
S. Huang, Q. Meng, A. Maminska, J. D. MacMicking, Cell-autonomous immunity by IFN-induced GBPs in animals and plants, Curr. Opin. Immunol., 60 (2019), 71–80. https://doi.org/10.1016/j.coi.2019.04.017 doi: 10.1016/j.coi.2019.04.017
![]() |
[3] |
K. Tretina, E. S. Park, A. Maminska, J. D. MacMicking, Interferon-induced guanylate-binding proteins: Guardians of host defense in health and disease, J. Exp. Med., 216 (2019), 482–500. https://doi.org/10.1084/jem.20182031 doi: 10.1084/jem.20182031
![]() |
[4] |
K. Lipnik, E. Naschberger, N. Gonin-Laurent, P. Kodajova, H. Petznek, S. Rungaldier, et al., Interferon gamma-induced human guanylate binding protein 1 inhibits mammary tumor growth in mice, Mol. Med., 16 (2010), 177–187. https://doi.org/10.2119/molmed.2009.00172 doi: 10.2119/molmed.2009.00172
![]() |
[5] |
P. Godoy, C. Cadenas, B. Hellwig, R. Marchan, J. Stewart, R. Reif, et al., Interferon-inducible guanylate binding protein (GBP2) is associated with better prognosis in breast cancer and indicates an efficient T cell response, Breast Cancer, 21 (2014), 491–499. https://doi.org/10.1007/s12282-012-0404-8 doi: 10.1007/s12282-012-0404-8
![]() |
[6] |
H. Xu, L. Sun, Y. Zheng, S. Yu, J. Ou-Yang, H. Han, et al., GBP3 promotes glioma cell proliferation via SQSTM1/p62-ERK1/2 axis, Biochem. Biophys. Res. Commun., 495 (2018), 446–453. https://doi.org/10.1016/j.bbrc.2017.11.050 doi: 10.1016/j.bbrc.2017.11.050
![]() |
[7] |
W. J. Chen, L. Xiong, L. Yang, L. J. Yang, L. Li, L. Huang, et al., Long Non-Coding RNA LINC01783 Promotes the Progression of Cervical Cancer by Sponging miR-199b-5p to Mediate GBP1 Expression, Cancer Manag. Res., 12 (2020), 363–373. https://doi.org/10.2147/CMAR.S230171 doi: 10.2147/CMAR.S230171
![]() |
[8] |
L. Cheng, L. Gou, T. Wei, J. Zhang, GBP1 promotes erlotinib resistance via PGK1activated EMT signaling in nonsmall cell lung cancer, Int. J. Oncol., 57 (2020), 858–870. https://doi.org/10.3892/ijo.2020.5086 doi: 10.3892/ijo.2020.5086
![]() |
[9] |
J. Song, Q. Y. Wei, GBP1 promotes non-small cell lung carcinoma malignancy and chemoresistance via activating the Wnt/beta-catenin signaling pathway, Eur. Rev. Med. Pharmacol. Sci., 24 (2020), 5465–5472. https://doi.org/10.26355/eurrev_202005_21331 doi: 10.26355/eurrev_202005_21331
![]() |
[10] |
X. Ji, H. Zhu, X. Dai, Y. Xi, Y. Sheng, C. Gao, et al., Overexpression of GBP1 predicts poor prognosis and promotes tumor growth in human glioblastoma multiforme, Cancer Biomark., 25 (2019), 275–290. https://doi.org/10.3233/CBM-171177 doi: 10.3233/CBM-171177
![]() |
[11] |
J. Zhao, X. Li, L. Liu, J. Cao, M. A. Goscinski, H. Fan, et al., Oncogenic Role of Guanylate Binding Protein 1 in Human Prostate Cancer, Front. Oncol., 9 (2019), 1494. https://doi.org/10.3389/fonc.2019.01494 doi: 10.3389/fonc.2019.01494
![]() |
[12] |
S. Yu, X. Yu, L. Sun, Y. Zheng, L. Chen, H. Xu, et al., GBP2 enhances glioblastoma invasion through Stat3/fibronectin pathway, Oncogene, 39 (2020), 5042–5055. https://doi.org/10.1038/s41388-020-1348-7 doi: 10.1038/s41388-020-1348-7
![]() |
[13] |
F. Rahvar, M. Salimi, H. Mozdarani, Plasma GBP2 promoter methylation is associated with advanced stages in breast cancer, Genet. Mol. Biol., 43 (2020), 20190230. https://doi.org/10.1590/1678-4685-GMB-2019-0230 doi: 10.1590/1678-4685-GMB-2019-0230
![]() |
[14] |
J. Zhang, Y. Zhang, W. Wu, F. Wang, X. Liu, G. Shui, et al., Guanylate-binding protein 2 regulates Drp1-mediated mitochondrial fission to suppress breast cancer cell invasion, Cell Death Dis., 8 (2017), 3151. https://doi.org/10.1038/cddis.2017.559 doi: 10.1038/cddis.2017.559
![]() |
[15] |
B. Liu, R. Huang, T. Fu, P. He, C. Du, W. Zhou, et al., GBP2 as a potential prognostic biomarker in pancreatic adenocarcinoma, PeerJ, 9 (2021), 11423. https://doi.org/10.7717/peerj.11423 doi: 10.7717/peerj.11423
![]() |
[16] |
J. Wang, H. Min, B. Hu, X. Xue, Y. Liu, Guanylate-binding protein-2 inhibits colorectal cancer cell growth and increases the sensitivity to paclitaxel of paclitaxel-resistant colorectal cancer cells by interfering Wnt signaling, J. Cell. Biochem., 121 (2020), 1250–1259. https://doi.org/10.1002/jcb.29358 doi: 10.1002/jcb.29358
![]() |
[17] |
H. Xu, J. Jin, Y. Chen, G. Wu, H. Zhu, Q. Wang, et al., GBP3 promotes glioblastoma resistance to temozolomide by enhancing DNA damage repair, Oncogene, 41 (2022), 3876–3885. https://doi.org/10.1038/s41388-022-02397-5 doi: 10.1038/s41388-022-02397-5
![]() |
[18] |
A. Blum, P. Wang, J. C. Zenklusen, SnapShot: TCGA-analyzed tumors, Cell, 173 (2018), 530. https://doi.org/10.1016/j.cell.2018.03.059 doi: 10.1016/j.cell.2018.03.059
![]() |
[19] |
G. T. Consortium, Human genomics. The Genotype-Tissue Expression (GTEx) pilot analysis: multitissue gene regulation in humans, Science, 348 (2015), 648–660. https://doi.org/10.1126/science.1262110 doi: 10.1126/science.1262110
![]() |
[20] |
T. Wu, Y. Dai, Tumor microenvironment and therapeutic response, Cancer Lett., 387 (2017), 61–68. https://doi.org/10.1016/j.canlet.2016.01.043 doi: 10.1016/j.canlet.2016.01.043
![]() |
[21] |
Z. Wang, L. Fu, J. Zhang, Y. Ge, C. Guo, R. Wang, et al., A comprehensive analysis of potential gastric cancer prognostic biomarker ITGBL1 associated with immune infiltration and epithelial-mesenchymal transition, Biomed. Eng. Online, 21 (2022), 30. https://doi.org/10.1186/s12938-022-00998-5 doi: 10.1186/s12938-022-00998-5
![]() |
[22] |
T. Jiang, T. Yang, Y. Chen, Y. Miao, Y. Xu, H. Jiang, et al., Emulating interactions between microorganisms and tumor microenvironment to develop cancer theranostics, Theranostics, 12 (2022), 2833–2859. https://doi.org/10.7150/thno.70719 doi: 10.7150/thno.70719
![]() |
[23] |
K. Li, Z. Zhang, Y. Mei, M. Li, Q. Yang, Q. Wu, et al., Targeting the innate immune system with nanoparticles for cancer immunotherapy, J. Mater. Chem. B, 10 (2022), 1709–1733. https://doi.org/10.1039/d1tb02818a doi: 10.1039/d1tb02818a
![]() |
[24] |
F. Petitprez, M. Meylan, A. de Reynies, C. Sautes-Fridman, W. H. Fridman, The tumor microenvironment in the response to immune checkpoint blockade therapies, Front. Immunol., 11 (2020), 784. https://doi.org/10.3389/fimmu.2020.00784 doi: 10.3389/fimmu.2020.00784
![]() |
[25] |
C. Guo, Z. Luo, D. Ismtula, X. Bi, H. Kong, Y. Wang, et al., TIGIT as a novel prognostic marker for immune infiltration in invasive breast cancer, Comb. Chem. High Throughput. Screen., 26 (2023), 639–651. https://doi.org/10.2174/1386207325666220629162823 doi: 10.2174/1386207325666220629162823
![]() |
[26] |
T. Li, J. Fan, B. Wang, N. Traugh, Q. Chen, J. S. Liu, et al., TIMER: A web server for comprehensive analysis of tumor-infiltrating immune cells, Cancer Res., 77 (2017), 108–110. https://doi.org/10.1158/0008-5472.CAN-17-0307 doi: 10.1158/0008-5472.CAN-17-0307
![]() |
[27] |
T. Li, J. Fu, Z. Zeng, D. Cohen, J. Li, Q. Chen, et al., TIMER2.0 for analysis of tumor-infiltrating immune cells, Nucleic Acids Res., 48 (2020), 509–514. https://doi.org/10.1093/nar/gkaa407 doi: 10.1093/nar/gkaa407
![]() |
[28] |
Z. Tang, B. Kang, C. Li, T. Chen, Z. Zhang, GEPIA2: an enhanced web server for large-scale expression profiling and interactive analysis, Nucleic Acids Res., 47 (2019), 556–560. https://doi.org/10.1093/nar/gkz430 doi: 10.1093/nar/gkz430
![]() |
[29] |
Z. Tang, C. Li, B. Kang, G. Gao, C. Li, Z. Zhang, GEPIA: a web server for cancer and normal gene expression profiling and interactive analyses, Nucleic Acids Res., 45 (2017), 98–102. https://doi.org/10.1093/nar/gkx247 doi: 10.1093/nar/gkx247
![]() |
[30] |
D. Szklarczyk, A. L. Gable, K. C. Nastou, D. Lyon, R. Kirsch, S. Pyysalo, et al., The STRING database in 2021: customizable protein-protein networks, and functional characterization of user-uploaded gene/measurement sets, Nucleic Acids Res., 49 (2021), 605–612. https://doi.org/10.1093/nar/gkaa1074 doi: 10.1093/nar/gkaa1074
![]() |
[31] |
G. D. Jr, B.T. Sherman, D. A. Hosack, J. Yang, W. Gao, H. C. Lane, et al., DAVID: database for annotation, visualization, and integrated discovery, Genome Biol., 4 (2003), 3. https://doi.org/10.1186/gb-2003-4-5-p3 doi: 10.1186/gb-2003-4-5-p3
![]() |
[32] |
E. Cerami, J. Gao, U. Dogrusoz, B. E. Gross, S. O. Sumer, B. A. Aksoy, et al., The cBio cancer genomics portal: an open platform for exploring multidimensional cancer genomics data, Cancer Discov., 2 (2012), 401–404. https://doi.org/10.1158/2159-8290.CD-12-0095 doi: 10.1158/2159-8290.CD-12-0095
![]() |
[33] |
Q. Wang, X. Wang, Q. Liang, S. Wang, L. Xiwen, F. Pan, et al., Distinct prognostic value of mRNA expression of guanylate-binding protein genes in skin cutaneous melanoma, Oncol. Lett., 15 (2018), 7914–7922. https://doi.org/10.3892/ol.2018.8306 doi: 10.3892/ol.2018.8306
![]() |
[34] |
S. Feng, D. E. Tuipulotu, A. Pandey, W. Jing, C. Shen, C. Ngo, et al., Pathogen-selective killing by guanylate-binding proteins as a molecular mechanism leading to inflammasome signaling, Nat. Commun., 13 (2022), 4395. https://doi.org/10.1038/s41467-022-32127-0 doi: 10.1038/s41467-022-32127-0
![]() |
[35] |
X. Li, D. Song, S. Su, X. He, F. Cao, C. Yang, et al., Critical role of guanylate binding protein 5 in tumor immune microenvironment and predictive value of immunotherapy response, Front. Genet., 13 (2022), 984615. https://doi.org/10.3389/fgene.2022.984615 doi: 10.3389/fgene.2022.984615
![]() |
[36] |
R. Kalluri, The biology and function of fibroblasts in cancer, Nat. Rev. Cancer, 16 (2016), 582–98. https://doi.org/10.1038/nrc.2016.73 doi: 10.1038/nrc.2016.73
![]() |
[37] |
H. Raskov, A. Orhan, J. P. Christensen, I. Gogenur, Cytotoxic CD8(+) T cells in cancer and cancer immunotherapy, Br. J. Cancer, 124 (2021), 359–367. https://doi.org/10.1038/s41416-020-01048-4. doi: 10.1038/s41416-020-01048-4
![]() |
[38] |
H. Rodriguez, R. Prados-Rosales, J. L. Lavin, M. Mazzone, J. Anguita, Editorial: Macrophage metabolism and immune responses, Front. Immunol., 11 (2020), 1078. https://doi.org/10.3389/fimmu.2020.01078 doi: 10.3389/fimmu.2020.01078
![]() |
[39] |
T. Liu, C. Han, S. Wang, P. Fang, Z. Ma, L. Xu, et al., Cancer-associated fibroblasts: an emerging target of anti-cancer immunotherapy, J. Hematol. Oncol., 12 (2019), 86. https://doi.org/10.1186/s13045-019-0770-1 doi: 10.1186/s13045-019-0770-1
![]() |
[40] |
W. Shen, P. A. Yao, W. Li, C. Gu, T. Gao, Y. Cao, et al., Cancer-associated fibroblast-targeted nanodrugs reshape colorectal tumor microenvironments to suppress tumor proliferation, metastasis and improve drug penetration, J. Mater. Chem. B, (2022). https://doi.org/10.1039/d2tb02253b doi: 10.1039/d2tb02253b
![]() |
[41] |
B. Farhood, M. Najafi, K. Mortezaee, CD8(+) cytotoxic T lymphocytes in cancer immunotherapy: A review, J. Cell. Physiol., 234 (2019), 8509–8521. https://doi.org/10.1002/jcp.27782 doi: 10.1002/jcp.27782
![]() |
[42] |
A. Mantovani, I. Barajon, C. Garlanda, IL-1 and IL-1 regulatory pathways in cancer progression and therapy, Immunol. Rev., 281 (2018), 57–61. https://doi.org/10.1111/imr.12614 doi: 10.1111/imr.12614
![]() |
[43] |
L. Ni, J. Lu, Interferon gamma in cancer immunotherapy, Cancer Med., 7 (2018), 4509–4516. https://doi.org/10.1002/cam4.1700 doi: 10.1002/cam4.1700
![]() |
[44] |
D. S. Chen, I. Mellman, Elements of cancer immunity and the cancer-immune set point, Nature, 541 (2017), 321–330. https://doi.org/10.1038/nature21349 doi: 10.1038/nature21349
![]() |
[45] |
J. Reiser, A. Banerjee, Effector, memory, and dysfunctional CD8(+) T cell fates in the antitumor immune response, J. Immunol. Res., 2016 (2016), 8941260. https://doi.org/10.1155/2016/8941260 doi: 10.1155/2016/8941260
![]() |
[46] |
C. Gerada, K. M. Ryan, Autophagy, the innate immune response and cancer, Mol. Oncol., 14 (2020), 1913–1929. https://doi.org/10.1002/1878-0261.12774 doi: 10.1002/1878-0261.12774
![]() |
[47] |
N. Karin, Chemokines and cancer: New immune checkpoints for cancer therapy, Curr. Opin. Immunol., 51 (2018), 140–145. https://doi.org/10.1016/j.coi.2018.03.004 doi: 10.1016/j.coi.2018.03.004
![]() |
[48] |
W. Zheng, D. R. E. Ranoa, X. Huang, Y. Hou, K. Yang, E. C. Poli, et al., RIG-I-like receptor LGP2 is required for tumor control by radiotherapy, Cancer Res., 80 (2020), 5633–5641. https://doi.org/10.1158/0008-5472.CAN-20-2324 doi: 10.1158/0008-5472.CAN-20-2324
![]() |
[49] |
C. Bourquin, A. Pommier, C. Hotz, Harnessing the immune system to fight cancer with Toll-like receptor and RIG-I-like receptor agonists, Pharmacol. Res., 154 (2020), 104192. https://doi.org/10.1016/j.phrs.2019.03.001 doi: 10.1016/j.phrs.2019.03.001
![]() |
[50] |
T. H. Tran, T. T. P. Tran, D. H. Truong, H. T. Nguyen, T. T. Pham, C. S. Yong, et al., Toll-like receptor-targeted particles: A paradigm to manipulate the tumor microenvironment for cancer immunotherapy, Acta Biomater., 94 (2019), 82–96. https://doi.org/10.1016/j.actbio.2019.05.043 doi: 10.1016/j.actbio.2019.05.043
![]() |
[51] |
P. Liu, Z. Lu, L. Liu, R. Li, Z. Liang, M. Shen, et al., NOD-like receptor signaling in inflammation-associated cancers: From functions to targeted therapies, Phytomedicine, 64 (2019), 152925. https://doi.org/10.1016/j.phymed.2019.152925 doi: 10.1016/j.phymed.2019.152925
![]() |
[52] |
J. Kwon, S. F. Bakhoum, The cytosolic DNA-sensing cGAS-STING pathway in cancer, Cancer Discov., 10 (2020), 26–39. https://doi.org/10.1158/2159-8290.CD-19-0761 doi: 10.1158/2159-8290.CD-19-0761
![]() |
[53] |
P. de Ron, S. Dremier, P. Winlow, A. Jenkins, E. Hanon, A. N. da Costa, Correlating behaviour and gene expression endpoints in the dopaminergic system after modafinil administration in mouse, Eur. Neuropsychopharmacol., 26 (2016), 729–740. https://doi.org/10.1016/j.euroneuro.2016.01.010 doi: 10.1016/j.euroneuro.2016.01.010
![]() |
1. | Qianqian Li, Fengde Chen, Lijuan Chen, Zhong Li, Dynamical Analysis of a Discrete Amensalism System with Michaelis–Menten Type Harvesting for the Second Species, 2024, 23, 1575-5460, 10.1007/s12346-024-01142-5 |
Conditions | Case | ||
0<a2<2 | sink | ||
e∗∗2<e2<e11+ka2 | 0<c<˜c | a2>2 | saddle |
a2=2 | non-hyperbolic | ||
0<a2<2 | saddle | ||
0<c<c1 | a2>2 | source | |
a2=2 | non-hyperbolic | ||
0<a2<2 | non-hyperbolic | ||
0<e2<e∗∗2 | c=c1 | a2>2 | non-hyperbolic |
a2=2 | non-hyperbolic | ||
0<a2<2 | sink | ||
c1<c<˜c | a2>2 | saddle | |
a2=2 | non-hyperbolic |
Values of bifurcation parameter k from the chaotic region | Stability interval ρ |
0.20 | 0<ρ<0.648648648648 |
0.25 | 0<ρ<0.680272108843 |
0.30 | 0<ρ<0.712328767123 |
0.35 | 0<ρ<0.744827586206 |
0.40 | 0<ρ<0.777777777777 |
0.45 | 0<ρ<0.811188811188 |
Conditions | Case | |
k>k∗ | The first species became extinct. | |
0<a2<2 | As k increases, the density of the first species decreases. | |
k∗∗<k<k∗ | When the first species is in chaos, k has a stabilizing effect. | |
k>k∗ | The first species became extinct. | |
a2>2 | 0<k<k∗ | Decrease in the density of the first species. |
Conditions | Case | ||
0<a2<2 | sink | ||
e∗∗2<e2<e11+ka2 | 0<c<˜c | a2>2 | saddle |
a2=2 | non-hyperbolic | ||
0<a2<2 | saddle | ||
0<c<c1 | a2>2 | source | |
a2=2 | non-hyperbolic | ||
0<a2<2 | non-hyperbolic | ||
0<e2<e∗∗2 | c=c1 | a2>2 | non-hyperbolic |
a2=2 | non-hyperbolic | ||
0<a2<2 | sink | ||
c1<c<˜c | a2>2 | saddle | |
a2=2 | non-hyperbolic |
Values of bifurcation parameter k from the chaotic region | Stability interval ρ |
0.20 | 0<ρ<0.648648648648 |
0.25 | 0<ρ<0.680272108843 |
0.30 | 0<ρ<0.712328767123 |
0.35 | 0<ρ<0.744827586206 |
0.40 | 0<ρ<0.777777777777 |
0.45 | 0<ρ<0.811188811188 |
Conditions | Case | |
k>k∗ | The first species became extinct. | |
0<a2<2 | As k increases, the density of the first species decreases. | |
k∗∗<k<k∗ | When the first species is in chaos, k has a stabilizing effect. | |
k>k∗ | The first species became extinct. | |
a2>2 | 0<k<k∗ | Decrease in the density of the first species. |