1. Introduction
Hyphanthia cunea (also called Fall webworm, H. cunea for short) is a kind of quarantine pest that poses a danger to forestry and agriculture. H. cunea originated in North America, and it was accidentally introduced to central Europe and eastern Asia in the early 1940s [1]. Now it has spread to more than 30 countries, including South Africa, Argentina, China, Belarus, etc., and has been classified as an invasive and quarantine insect by organizations such as APPPC, COSA VE and EAEU EPPO (from Global Data Base: https://gd.eppo.int/taxon/HYPHCU/distribution). H. cunea has more than 600 recorded host plant species, which are widely distributed in several countries such as the United States, Japan, Korea, and China [2]. Due to its wide distribution across various host plants, H. cunea causes substantial economic losses globally. With the onset of global warming, H. cunea is increasingly spreading toward higher latitudes and elevations. Since it was first discovered in 1979 in Dandong City, Liaoning Province, China [3], the pest has spread to the Yangtze River basin, involving more than 10 provinces (cities and autonomous regions).
Both larvae and adults of the H. cunea will gnaw on leaves and even trunks. The species is highly reproductive, spreads rapidly, and causes significant damage to forests [4] [5]. H. cunea can threaten the survival of local species, diminish biodiversity and endanger ecological security. Figure 1 shows the condition of trees infected by the H. cunea. The leaves of the trees become transparent after being eaten by H. cunea (as in Figure 1(a) and Figure 1(b)), and the trunks of the trees turn yellow and wilt after being eaten (as in Figure 1(c) and Figure 1(d), Figure 1(c) is a partially enlarged image of Figure 1(d)) (The four images in Figure 1 are from: https://baike.baidu.com).
Figure 1. Leaves and trunks infected with the H. cunea.
The infestation of H. cunea mainly has the following control methods:
(1) Physical control method: This method primarily involves manual interventions such as cutting net screens, capturing adults, and digging pupae manually.
(2) Chemical pharmaceutical control method: This method mainly involves spraying susceptible trees with pesticides.
(3) Biological control method: This method involves releasing natural enemies of H. cunea, with the most effective and widely used being C. cunea.
The first two methods are effective, however, they either require a lot of manpower and material resources or can potentially damage the ecosystem. Biological control is the most environmentally friendly and economical method. In order to explore the biological control method of the H. cunea, Yang et al. [3] investigated the parasitic natural enemies of the H. cunea continuously from 1984 to 1986, and finally found a new natural enemy in the pupae of the H. cunea in Shanxi province. C. cunea is a dominant parasitic wasp with high parasitism and high fecundity. It is sensitive to lepidopteran pests such as H. cunea. C. cunea can insert its ovipositor into the pupae of insect pests, where it develops and grows, consuming all the nutrients and eventually causing the demise of H. cunea. H. cunea progresses through four life stages: egg, larva, pupa, and adult. The female wasps can lay eggs on the same day after eclosion under suitable temperature conditions. Female and male wasps mate within the host pupae and then emerge to seek a new host pupa for egg-laying. The release rate of C. cunea is determined by the density of H. cunea [6]. Figure 2 shows the reproductive mechanism of the coexistence of the H. cunea and the C. cunea in nature, with the red arrow indicating the artificial release of C. cunea. The left circle of Figure 2 shows the reproduction mechanism of the H. cunea, and the right circle shows the reproduction mechanism of the C. cunea. The intersection of the two circles is the H. cunea pupa.
![]()
Figure 2. Reproductive mechanism of H. cunea and C. cunea.
In recent years, some scholars have used ecological niche models (ENMs) to combine H. cunea occurrence records with environmental data. By creating relevant models to predict their ecological demand and potential geographic distribution. These experiments show that incorporating global data can substantially increase the forecasting precision of ENMs. This will provide valuable insights for the control of wedge nematode disease in China [7]. Several studies have investigated how temperature and season affect the behaviour of H. cunea populations and have created population dynamics models for its different life stages [8]. In addition, single population models with time delay were created [2]. Therefore, it is valuable to describe the dynamics of H. cunea and C. cunea by applying mathematical models.
In some sense, the richer dynamic properties of the model may be induced by time delay. Martin and Ruan [9] [10] proposed three types of time delay, these terms manifest within the prey-specific growth period, the predator response component of the predator equation, and the interaction component of the predator equation. Jiang et al. [11] [12] developed a predator-prey model incorporating time delay and explored a variety of dynamic properties associated with it. Time delay is crucial for the emergence of oscillatory and periodic behaviours. Considering that C. cunea can inhibit the growth rate of H. cunea, but this inhibition is not instantaneous. Therefore, the inclusion of time delay produces richer dynamic properties. The research on current time delay models has been a hot topic in mathematics, physics, biology and other fields. Furthermore, the introduction of nonlocal competition and its impact on model stability has been a subject of intense discussion in the literature over the past few decades. The works of Britton et al. [13] [14] have made significant contributions to the advancement of nonlocal competition. In their respective papers, they considered nonlocal competition and incorporated spatial convolution integrals into single species models. Britton et al. suggested that resource utilization in a given area is influenced by two factors: the density of the local population and the weighted average populations in surrounding areas. Song et al. [15]-[18] incorporated nonlocal competition into a predator-prey model with time delay and discovered that the interplay between nonlocal competition and time delay causes instability in the equilibrium.
The structure of the paper is as follows: Section 2 describes the process of establishing the mathematical model. Section 3 investigates the existence and stability of the positive equilibria and the existence of Hopf and Turing bifurcations near the equilibria. In Section 4, we utilize the multiple time scales method to derive the normal form of Hopf bifurcation for the model incorporating nonlocal competition. Section 5 presents numerical simulations that demonstrate our main results. Section 6 provides the conclusion.
2. Mathematical Modeling
In this section, we show the mathematical modeling process and explain it. Firstly, as noted in Ref. [19], the Lotka-Volterra and Nicholson-Bailey models were two primary frameworks traditionally employed to study predator-prey and parasitic-host interrelations. The Lotka-Volterra model results in neutrally stable cycles, while the Nicholson-Bailey system leads to diverging cycles and eventual extinction. Specifically, the Lotka-Volterra model can be expressed as follows:
(1)
here,
and
represent the prey and predator populations, respectively. The parameters
and
denote the highest per capita growth rate and environmental capacity of prey, respectively.
represents the capture rate of prey.
represents the efficiency of converting ingested prey into new predators.
represents the per capita mortality rate or the per capita nutritional needs for sustaining and replacing predators. Taking into account both biological and physical control, we establish the H. cunea-C. cunea model as follows:
(2)
where
and
have the same meanings as in model (1).
and
represent the population density of H. cunea and C. cunea, respectively.
represents the probability of the H. cunea pupa being parasitized by the C. cunea.
represents the proportionality coefficient.
represents the mortality rate of H. cunea caused by physical control methods.
represents the natural mortality rate of C. cunea.
represents the probability of successful eclosion of the parasitized pupae.
represents the release rate of C. cunea. This paper intends to study the method of reducing the density of H. cunea in more environmentally friendly ways. Therefore, the model (2) does not include the death rate of H. cunea due to pesticides. Reducing the population of H. cunea can be achieved by regularly releasing a quantified amount of C. cunea.
We also take into account the time delay, as the eclosion of H. cunea pupae is not instantaneous, regardless of parasitization. Since the consumption of environmental resources is affected by the flow of H. cunea, thus, the competition is not local. A plethora of researchers incorporated nonlocal competition by substituting the original
with
, where , which reflects intraspecific competition within prey populations. Note that
with
is a positive parameter that measures the length of the habitat, and
represents the habitat volume of H. cunea. We assume that
as
, and we study the model (3) with Neumann boundary condition to facilitate the study of nonlocal competition. To sum up, we develop a H. cunea-C. cunea model with both nonlocal competition and time delay as follows:
(3)
where
,
represent the diffusion coefficients of H. cunea and C. cunea, respectively. The other parameters
,
,
,
,
,
,
and
represent the same meanings as the parameters in model (2), and all of them are non-negative.
Next, we will analyze the dynamic properties of model (3).
3. Existence and Stability of Equilibria and Existence of
Bifurcations
In this section, we will consider the existence and stability of the equilibria for model (3), as well as the existence of Hopf and Turing bifurcations near the equilibria.
We know that model (3) has a trivial equilibrium
, and we make the following hypothesis:
(H1)
if (H1) holds, then the model (3) has a positive equilibrium
, where
with
, it is obvious that
, and
when
.
Remark 1. The death rates from physical insect control methods are generally low, such as traps, barriers, electric shocks, etc., which are often non-lethal. The objective of physical control typically aims to diminish insect populations rather than inducing immediate widespread death.
3.1. Existence of Hopf Bifurcation
Next, we analyze the stability of the equilibria for the model (3), and model (3) can be shown as follows:
(4)
where
with
,
,
,
,
,
. For trivial equilibrium
, the characteristic equation corresponding to model (3) at
has two roots:
,
. Therefore, the trivial equilibrium
is a saddle point.
The characteristic equation corresponding to model (3) at
is as follows:
(5)
where
, when
,
,
,
. When
,
,
,
.
When
, characteristic Eq. (5) becomes
Next, we make the following hypothesis:
(H2)
clearly,
, and
. This is because both
and
are monotonically increasing for
. To sum up, if (H2) holds,
and
hold for any
.
Based on the above analysis, we give the following theorem:
Theorem 1. If hypothesis (H1) holds,
is a positive equilibrium, and
is a saddle point. If hypotheses (H1) and (H2) hold, the positive equilibrium
of model (3) for
is locally asymptotically stable for any integer
.
Next, we explore the existence of Hopf bifurcation near the equilibrium
for
. Let
are the characteristic roots of Eq. (5). Separating the real part and imaginary part, we obtain
(6)
that is,
(7)
letting
, then we get
(8)
By Eq. (6), we can get
then for
and
, we get
(9)
it is clear that
monotonically increases with
. By calculation, the transversality condition
according to Eq. (8), it is clearly that
for any
. Thus
holds, then we denote
(10)
where
is given in Eq. (9).
Denote that
and
are sets of integers, which may be finite or infinite,
where
. Combining the analysis above, we derive the following theorem.
Theorem 2. Assuming that hypotheses (H1) and (H2) are satisfied, and
is given in Eq. (9). Thus, we have following conclusions:
(1) If
and
, then equilibrium
is locally asymptotically stable for
.
(2) If
and
, Eq. (8) has a unique positive root
for some
, then the equilibrium
of model (3) is locally asymptotically stable for
and becomes unstable for
, where
. Additionally, model (3) undergoes
-mode Hopf bifurcation near the equilibrium
for
with
and
.
(3) If
and
, Eq. (8) has two positive roots
for some
, then the equilibrium
of the model (3) is locally asymptotically stable for
, and there may be a switch of stability for
in the model (3), where
. Additionally, model (3) undergoes
-mode Hopf bifurcation near the equilibrium
for
with
and
.
(4) If
and
, Eq. (8) has positive roots
and
for some
,
, respectively. Then the equilibrium
of the model (3) is locally asymptotically stable for
, and there may be a switch of stability for
in the model (3), where
. Additionally, model (3) undergoes
-mode Hopf bifurcation near the equilibrium
for
with
, or
and
.
3.2. Existence of Turing Bifurcation
In this subsection, we analyze the conditions for the existence of Turing bifurcation. Suppose
is the root of characteristic Eq. (5), then we get expression:
(11)
where
due to
and
are positive.
Letting
, by simple calculation from Eq. (11), we get
According to its biological significance that
and
, which means
and
, then we have
Then we making the following hypothesis:
(H3)
We denote that
, and we
have the following theorem.
Theorem 3. If (H3) is true and a positive integer
can be found, then the model (3) undergoes the
-mode Turing bifurcation. If
, then model (3) does not undergo Turing bifurcation.
4. Stability and Direction of Hopf Bifurcation
In this section, we derive the normal form of the Hopf bifurcation for the model (3) by using the multiple time scales method as referenced in ref. [20]-[22]. When
, the characteristic equation Eq. (5) has a pair of pure imaginary roots
, at which model (3) undergoes
,
or
-mode Hopf bifurcation at equilibrium
. In the subsequent derivation, we simplify notation by using
instead of
,
or
. We consider the time delay
serves as the bifurcation parameter in this paper, where
, with
being the critical value for Hopf bifurcation. The parameter
is a dimensionless scaling factor, and
is a perturbation parameter. We also introduce the transformations:
,
,
, then model (3) is transformed into
(12)
Let
be the eigenvector of the linear operator corresponding to the eigenvalue
of the linearized system of Eq. (12) for equilibrium
, and let
be the normalized eigenvector of the adjoint operator of the linear operator corresponding to the eigenvalue
, satisfying
. By calculation, it can be found
(13)
where
, and
.
According to the multiple scales method, we can assume the solution of Eq. (12) as follows:
(14)
where
, specially,
(15)
By substituting (14) and (15) into (12), we compare the coefficients of
and
on either side of the equation, resulting in the following expressions:
(16)
(17)
(18)
Referring to the notations in Refs. [20] [21], then Eq. (16) has the solution with following form:
(19)
Therefore, by substituting Eq. (19) into the right part of Eq. (17), then we define the coefficient vector of
as
. Therefore, according to the solvability condition, we have
. Then we can solve
, namely,
(20)
where
with
Suppose that the solution of (17) satisfies following form:
(21)
we substitute (21) into the left end of (17), it can be obtained by the method of undetermined coefficients, where
with
and
Next, substitute (19) and (21) into (18), then compare the coefficients, and denoting the coefficient vector of
as
. Therefore, according to the solvability condition, we have
. Specially,
is a perturbation parameter, and
has minimal influence of small unfolding parameters. Thus, we have
(22)
where
with
and
We drive the normal form of Hopf bifurcation for model (3) reduce on the center manifold as follows:
(23)
where
and
are defined by (20) and (22), respectively. Substituting
into the Eq. (23), and we derive the amplitude equation on the center manifold as follows:
then we state the following theorem.
Theorem 4. If the condition
holds, the model (3) generates periodic solutions near the equilibrium
. If
, the bifurcating periodic solutions are unstable (stable). Additionally, the direction of bifurcating periodic solution is forward (backward) if
(
).
5. Numerical Simulations
In this section, we first analyze the parameter ranges, then simulations are performed numerically.
5.1. Selection of Parameter Values
(1) Releasing rate
and natural mortality
of C. cunea.
In the process of biological control of H. cunea, the release rate of C. cunea depends on the density of H. cunea. According to Ref. [6], the ratio of the C. cunea to the H. cunea is 3:1 in areas with more pests. Thus, we first consider the case when
, and we also consider what happens when
takes different values. According to Ref. [23], the mortality rate varies across different life stages of insects and may range from 0.1 to 0.9. Consequently, we set the natural mortality rate of C. cunea as
in our paper.
(2) Self-diffusion coefficients
and
.
Considering the distinct life stages, it's evident that the egg and pupa stages are relatively stationary. Consequently, the self-diffusion coefficient
for H. cunea can be conservatively estimated at 0.0003 due to limited mobility during these stages. C. cunea lays eggs in the pupae of H. cunea, thus, it moves with the movement of the H. cunea pupae but slightly faster. Therefore, we set its diffusion coefficient to
.
(3) Net birth rate
and physical control fatality rate
of H. cunea.
First of all,
where
is birth rate of H. cunea.
is natural death rate of H. cunea, and
is death rate of predatory. According to the life table in Ref. [24], it showed mortality records for H. cunea across all stages due to various causes, and the range of H. cunea birth rate is
, thus, we select
. The range of H. cunea natural mortality rate is
, and we select
. The range of death rate of predatory is
, and we select
. Therefore,
. Moreover, we select
according to the analysis of Remark 1.
(4) Parasitism rate
and survival rate
of C. cunea.
In Ref. [25], the experiment of rearing C. cunea with Tussah (Silkworm) pupae was carried out by three-cut method and inoculating method, and the range of parasitism rate is
. The range of survival rate of H. cunea is
. In our paper, we select
and
, this is because selecting intermediate high values for
and
can provide a more robust and conservative estimate in experimental design. This helps ensure the reliability of experimental results and offers higher feasibility and effectiveness in practical applications.
(5) Other dimensionless parameters
and
.
We may select
and
for the environmental capacity and proportionality coefficient of H. cunea.
To sum up the analysis, we select,
,
,
,
,
,
,
,
,
,
.
5.2. Simulations
In this subsection, we will show the existence of various spatiotemporal patterns for model (2) and model (3).
5.2.1. Non-spatial Model (2)
The hypotheses (H1) and (H2) hold when the parameters
,
,
,
,
,
,
,
,
are fixed and
. According to Theorem 1, the equilibrium
is locally asymptotically stable for
and
. Figure 3 shows the change in the value of
when
. Fig. 3 shows that as the release rate of C. cunea increases, and then the value of
gradually decreases.
Biological interpretation 1:
Figure 3 shows that the population of the H. cunea reaches highest value when
. This is attributed to the lack of effective natural enemies, which led to the rapid growth of the population. Furthermore, it can be observed that the populations of H. cunea and C. cunea tend to stabilize when
. This is because they can quickly find and parasitize most of the H. cunea pupae when the amount of C. cunea is large enough, thereby inhibiting the increase in the H. cunea population. At the moment, the continuous parasitic action of the C. cunea keeps the H. cunea population at a low and stable level, rather than continuing to decline to near extinction.
Figure 3. The change of the equilibrium
as
increase.
Case 1:
:
Considering the case with biological control, that is,
, the equilibrium
. This set of parameters satisfies hypotheses (H1) and (H2) according to Theorem 1, and the positive equilibrium
of model (2) is locally asymptotically stable (see Figure 4). Figure 4(a) shows that the vector field distribution of equilibria for non-spatial model (2) when
. It is evident that
is a saddle point, and
is locally asymptotically stable. Figure 4(b) shows the locally asymptotic stability of the equilibrium
.
(a) (b)
Figure 4. The stability of equilibrium
of model (3) when
.
Case 2:
:
Next, we consider the case of biological control is absent, that is,
, the equilibrium
. This set of parameters makes
, then model (2) may generate periodic solution near the positive equilibrium
. Figure 5(a) shows that the vector field distribution of equilibria for non-spatial model (2) when
, and it can be seen that the
is a saddle point and the model (2) generates a stable periodic solution near the equilibrium
. Figure 5(b) shows periodic solution near the equilibrium
.
(a) (b)
Figure 5. The vector field near the equilibria
and
, and periodic solution near the equilibrium
of model (2) when
.
Biological interpretation 2:
(1) Figure 4 reflects the effectiveness of the strategy of controlling the H. cunea population through the artificial release of C. cunea. Initially, both the H. cunea and C. cunea populations show significant oscillations, then gradually stabilize. The H. cunea population decreases rapidly, while the C. cunea population increases and gradually stabilizes. This indicates that the artificial release of C. cunea can effectively control the H. cunea population.
(2) Figure 5 indicates that both H. cunea and C. cunea experience periodic natural eruptions when
. This erupt periodically is the result of the complex interactions between the H. cunea and C. cunea, reflecting the population regulation mechanisms in the natural ecosystem.
5.2.2. Spatial Model (3)
Case 1:
:
The set of parameters
,
,
,
,
,
,
,
,
,
satisfies hypotheses (H1) and (H2),
,
and
, then according to Theorem 2 and Eq. (10), we get the critical value
. Therefore, the positive equilibrium
is locally asymptotically stable for
and unstable for
. Figure 6 shows the asymptotic stability of the equilibrium
when
.
Figure 6. For
, when
,
of model (3) is locally asymptotically stable for
.
By calculating Eq. (20) and Eq. (22), we obtain
according to Theorem 4, the bifurcating periodic solution is stable and forward. The model (3) will generate inhomogeneous periodic solutions near the equilibrium
of model (3) for
(see Figure 7).
Figure 7. For
, the inhomogeneous stable and forward periodic solutions for model (3) near the equilibrium
when
.
We also consider the dynamics for the model (3) without nonlocal competition. Figure 8 shows that model (3) only generates stable homogeneous periodic solutions near the equilibrium
for
if there is no nonlocal competition.
Figure 8. For
, the homogeneous stable and forward periodic solutions for model (3) without nonlocal competition near the equilibrium
when
.
Case 2:
:
Considering spatial model (3) without biological control, that is,
. This set of parameters satisfies hypotheses (H1) and (H2). Similarly, we get the critical value
, thus, the equilibrium
is locally asymptotically stable for
and becomes unstable for
. By calculation,
,
,
and
,
. According to Theorem 4, the bifurcating periodic solution is stable and forward, then model (3) with
will generate inhomogeneous periodic solutions near
for
(see Figure 9).
Figure 9. For
, the inhomogeneous stable and forward periodic solutions for model (3) near the equilibrium
when
.
Biological interpretation 3:
(1) Figure 6 indicates that releasing C. cunea within the time delay
can effectively control the population of H. cunea. The H. cunea population is gradually controlled both temporally and spatially, and tends to stabilize.
(2) Figure 7 shows that releasing C. cunea after the critical value
will lead to periodic eruptions and instability in the H. cunea population. This means the H. cunea population cannot be effectively controlled and continues to pose a threat to the environment.
(3) By comparing Figure 7 and Figure 8, we find that the model (3) with nonlocal competition produces richer dynamics, which means the population dynamics are more similar to the behavior of the real ecosystem. In this way, the dynamic changes of populations in nature can be better simulated and understood. The phenomenon in Figure 8 may be due to the fact that H. cunea and C. cunea gather at the boundary when they fly to the boundary. When the density of H. cunea decreases, C. cunea flies in the opposite direction. Eventually, it causes large periodic oscillations on both sides, with equal numbers flying in opposite directions.
(4) By comparing Figure 7 and Figure 9, we find that when
, the stability interval of model (3) becomes smaller, and the amplitude of the inhomogeneous periodic solution becomes bigger. In summary, without biological control, the equilibrium
becomes unstable more rapidly and the model is prone to periodic eruptions. This highlights that the artificial release of C. cunea is a critical measure for maintaining ecological balance and effectively controlling the population of the H. cunea. It also reflects the inadequacy of natural populations regulation mechanisms.
Case 3:
:
By calculation, the hypotheses (H1) and (H2) hold when
,
,
,
,
,
,
,
,
, and
. When
, we get the critical value
at
, and when
, we get the critical value
at
. Figure 10 shows that
increases gradually with the increases of
.
Figure 10. The value of
with
gradually increasing.
Biological interpretation 4:
The nonlocal competition influences model (3) when
but does not impact model (3) when
. Therefore, nonlocal competition has little effect on the model when the population of C. cunea is large enough to cover every area of the forest, leaving no escape for H. cunea.
Suggestions:
(1) According to the Biological interpretation 1, the amount of releasing C. cunea should be controlled. Releasing too much C. cunea will put pressure on the environment and cause unnecessary economic losses.
(2) According to Biological interpretation 2 and 3, it is necessary to select the appropriate time and place for releasing C. cunea. Releasing C. cunea can effectively control the density of H. cunea before most pupae emerge. At the same time, the deployment of releasing C. cunea is optimized to ensure their migration to areas with high H. cunea concentrations, thereby improving parasitic efficiency on H. cunea pupae. The migration patterns of H. cunea are regularly monitored to ensure effective targeting and control of the H. cunea population in different spatial locations.
(3) According to Biological interpretation 4, the influence of nonlocal competition can be ignored in the process of analyzing the models when
.
(4) Based on the above analysis, our findings demonstrate that the population of artificially released C. cunea only requires reaching threefold the population size of H. cunea to achieve rapid suppression of H. cunea growth. This optimal ratio suggests that excessive release of C. cunea is unnecessary in practical applications, thereby effectively conserving both labor resources and economic expenditures.
6. Conclusions
In this paper, we studied a reaction-diffusion model of H. cunea and C. cunea with both time delay and nonlocal competition for biological control. We analyzed the existence and stability of the equilibria and the existence of Hopf and Turing bifurcations near the equilibria. We used the multiple time scales method to derive the normal form of the Hopf bifurcation reduced on the center manifold. Finally, a suitable set of parameters was selected for conducting numerical simulations. Spatial and non-spatial models were equally important aspects of our research. In the non-spatial model, we compared and analyzed the phenomena resulting from the with and without biological control. In the spatial model, we compared models with and without nonlocal competition, as well as with and without biological control. We found that the nonlocal competition can induce stable inhomogeneous periodic solutions of Hopf bifurcating. In the model without biological control, the critical time delay of the model decreased and generated periodic solutions with large amplitudes.
Results indicated that nonlocal competition significantly altered the spatial distribution and oscillation patterns of populations within the ecosystem. This phenomenon emphasised the need for geographically and environmentally sensitive biological control strategies. Effective biological control measures, particularly through the timely release of parasitic natural enemies, provided sustainable management and protection of H. cunea populations and ecosystems. The theoretical framework of the model, encompassing bifurcation analysis and the stability of periodic solutions, offered valuable theoretical insights for analyzing other invasive species-natural enemy systems. Future studies could validate the applicability of similar systems by altering the functional response structure of the model or adjusting the dispersal mechanism.
Acknowledgements
This study was funded by Fundamental Research Funds for the Central Universities (Grant No. HFW230600022) and Heilongjiang Provincial Natural Science Foundation of China (Grant No. LH2024A001).
NOTES
*Corresponding author.