1 Introduction

Since the pioneering work of Speirs and Gurney on the “drift paradox” (Speirs and Gurney 2001), studying population dynamics in advective environments (such as rivers) has become an active research topic, both empirically as well as theoretically (Fagan 2002; Grant et al. 2007; Huang et al. 2016; Jin et al. 2014, 2019; Jin and Lewis 2011, 2012; Lam et al. 2016; Lutscher et al. 2006, 2007, 2005; Pachepsky et al. 2015; Vasilyeva and Lutscher 2011). Most mathematical models in spatial ecology assume that individuals adopt random movement, i.e., the transition probability in all directions are the same. For the organisms in advective environments, they are also subject to the passive, uni-directional drift. Such passive drift may push the organisms to the downstream where the environments could become unfavorable. From the mathematical point of view, the addition of drift makes the differential operators under consideration non-symmetric and thus brings new challenges to the stability analysis, especially for those models for interacting species (Lou et al. 2016, 2019; Vasilyeva 2017; Vasilyeva and Lutscher 2012a, b; Wang and Shi 2019; Wang et al. 2019; Wang and Shi 2020; Zhao and Zhou 2016; Zhou 2016). Almost all of these studies assume that the underlying habitat is an interval in the real line, in order to simplify the mathematical analysis. In contrast, there are rather few studies on the population dynamics in river networks, and they are mostly restricted to the case of a single species (Jin et al. 2019; Ramirez 2012; Samia et al. 2015; Sarhad et al. 2014; Vasilyeva 2019).

One important topic in spatial ecology is the evolution of dispersal. The seminal work of Hastings shows that random dispersal is selected against in spatially heterogeneous but temporally constant environments (Hastings 1983; Dockery et al. 1998), while in spatially and temporally varying environments large dispersal rate can be selected (Hutson et al. 2001; McPeek and Holt 1992). See the review article (Cosner 2014) and the references therein. The evolution of dispersal in advective and continuous habitats has been recently considered: when the carrying capacity is spatially heterogeneous and the drift rates are constants, some intermediate dispersal rate could be selected; see (Golubitsky et al. 2017; Lam et al. 2014). However, for a homogeneous environment where both carrying capacity and drift rates are spatially uniform, it was shown that when the loss at the downstream is not significant, the faster dispersal rate is favored (Lou and Lutscher 2014; Lou and Zhou 2015); see (Hao et al. 2021) for more recent progress. Again, these studies assume that the habitat is a finite interval.

Many of the above work employ the conceptual framework of adaptive dynamics theory (Dieckmann and Law 1996; Geritz et al. 1998). A central idea of adaptive dynamics theory is the evolutionarily stable strategy (ESS) (Maynard-Smith and Price 1973). A strategy is called a global ESS if the resident species adopting such a strategy cannot be invaded by any rare mutant species using different strategy. Another important concept is the convergence stable strategy (CvSS). A strategy is said to be a global CvSS if the mutant species is always able to invade a resident species when the mutant strategy is closer to the CvSS than the resident strategy. Local ESS and CvSS can be similarly defined and interpreted.

Our aim is to study the evolution of dispersal in discrete, advective environments using the conceptual framework of adaptive dynamics theory. Recently, the authors proposed in (Jiang et al. 2020) to study the dynamics of two competing species in three-patch models with different network topology, and to investigate how the topology of directed river network modules may affect the evolution of dispersal. To be specific, we considered the following three types of river network modules in (Jiang et al. 2020):

Fig. 1
figure 1

Three river network modules with different topology: The two-way blue arrows represent the dispersal of species between connected patches, the one-way red arrows represent the uni-directional drift. The parameters dD are dispersal rates for the two competing species, and the parameters \(q_1,q_2\) are drift rates from an upstream patch to a downstream patch (Color figure online)

In Fig. 1a–c we assume that patch 1 is upstream, patch 3 is downstream, and patch 2 is either upstream, or middle stream, or downstream. In Jiang et al. (2020) the carrying capacity of three patches is assumed to be different and the drift rates are assumed to be equal. The main findings in Jiang et al. (2020) are summarized as follows: when the drift rate is small, for all three models the mutant species can invade when rare if and only if it is the slower disperser. However, when the drift rate is large, Models I and II predict that the faster disperser wins, while Model III predicts that fast and slow dispersers may coexist, and that there exists one intermediate strategy which is evolutionarily singular. For the intermediate range of drift, Models I and II predict the existence of one singular strategy, which may or may not be evolutionarily stable, depending upon the topology of modules, while Model III predicts singular strategy may not exist and the faster disperser wins the competition.

The rest of this paper is organized as follows: In Sect. 2 we state the main results for three-patch models. In Sect. 3 we draw the main conclusions and also provide a single framework to unify our main results. In Sect. 4 we present the numerical simulations of some 4-patch models and discuss some predictions on n-patch models. The proofs of the main results for Model I to III are postponed to the Appendices.

2 Main Results for Three-patch Models

In this paper, we assume that the drift rates could be different but the carrying capacity is the same in all three patches. As in Jiang et al. (2020), in all models the two competing species are assumed to be identical except for their dispersal rates.

Our main goal in this paper is to illustrate the effects of varying drift rates and network topology on the evolution of dispersal.

The main findings can briefly be summarized as follows:

  • If all drift rates are identical, then the faster dispersal rate is selected across all three-patch models in which the drift network do not form a closed cycle.

  • For general drift rates, infinite diffusion rate is a local CvSS for all three models.

  • For Models II and III, when a singular strategy (that is neither zero nor infinity) exists, it is not a local CvSS (Numerical simulation suggests that it is not an ESS either).

  • For Models II and III, when bi-stability occurs, it is possible for two competing populations with different dispersal rates to coexist, by varying the drift rates between patches.

2.1 Main Results of Model I

In Model I, the species in patches 1 and 2 are washed down to patch 3 by drift with rates \(q_1, q_2\), respectively. Two competing populations can disperse freely between the upstream patches and the downstream patch, with respective rates dD. The two upstream patches, however, are not directly connected. The diagram of Model I is shown in Fig. 1a. The dynamics of two competing populations in this river module is described by the following system of ODEs:

$$\begin{aligned} \left\{ \begin{array}{l} \frac{du_1}{dt}=d(u_3-u_1)-q_1u_1+u_1(1-\frac{u_1+v_1}{k})\\ \frac{du_2}{dt}=d(u_3-u_2)-q_2u_2+u_2(1-\frac{u_2+v_2}{k})\\ \frac{du_3}{dt}=d(u_1+u_2-2u_3)+q_1u_1+q_2u_2+u_3(1-\frac{u_3+v_3}{k})\\ \frac{dv_1}{dt}=D(v_3-v_1)-q_1v_1+v_1(1-\frac{u_1+v_1}{k})\\ \frac{dv_2}{dt}=D(v_3-v_2)-q_2v_2+v_2(1-\frac{u_2+v_2}{k})\\ \frac{dv_3}{dt}=D(v_1+v_2-2v_3)+q_1v_1+q_2v_2+v_3(1-\frac{u_3+v_3}{k})\\ u_i(0)=u_{i0},\quad v_i(0)=v_{i0},\quad i=1,2,3.\\ \end{array} \right. \end{aligned}$$
(1)

Here \(u_i(t),v_i(t)\) \((i=1,2,3)\) denote the numbers of individuals of the respective species at time t in patch i.

The parameter k is the carrying capacity for all patches. For the sake of simplicity, the intrinsic growth rates are assumed to be equal to 1. The initial data of \(u_i\) and \(v_i\) are assumed to be positive for the rest of the paper so that \(u_i(t), v_i(t)\) are positive functions of time \(t>0\).

It can be shown that system (1) has a unique semi-trivial steady state of the form \((U^*,0)=(U^*_1,U^*_2,U^*_3,0,0,0)\), where \(U^*_i>0\) for \(i=1,2,3\).

Theorem 1

For any \(q_1\ge 0\), \(q_2\ge 0\) and \(q_1 +q_2>0\), if \(d>D\), then \((U^*,0)\) is globally asymptotically stable among all solutions of (1) with positive initial data.

This result implies that the faster dispersal is always selected for Model I, provided that the carrying capacity is uniform in the habitat, and the conclusion is independent of the drift rates. The underlying biological intuition is that a single population at equilibrium (i.e., resident) is undermatching the resources in at least one of the two upstream patches and it is always overmatching the resource in the downstream patch; i.e., the downstream patch is always a sink and at least one of the upstream patches is a source. If a mutant with small diffusion rate is introduced, its individuals in the upstream patches will be washed to the downstream patch, where the mutant cannot invade when rare as the downstream patch is a sink. Hence, small diffusion rate is selected against. In contrast, faster diffusion can counterbalance the drift by keeping more mutant individuals in upstream patches, one of which is a source patch, and thus help the mutant populations establish in this upstream source patch.

2.2 Main Results of Model II

Model II assumes that individuals in patch 1 are transported to patch 2 by drift with rate \(q_1\), and individuals in patch 2 are transported to patch 3 by drift with rate \(q_2\). Individuals can also move between patches i and \(i+1\) for \(i=1,2\); see Fig. 1b. The dynamics of two competing species in this network module is governed by the following ODE system:

$$\begin{aligned} \left\{ \begin{array}{l} \frac{du_1}{dt}=d(u_2-u_1)-q_1u_1+u_1(1-\frac{u_1+v_1}{k})\\ \frac{du_2}{dt}=d(u_1+u_3-2u_2)+q_1u_1-q_2u_2+u_2(1-\frac{u_2+v_2}{k})\\ \frac{du_3}{dt}=d(u_2-u_3)+q_2u_2+u_3(1-\frac{u_3+v_3}{k})\\ \frac{dv_1}{dt}=D(v_2-v_1)-q_1v_1+v_1(1-\frac{u_1+v_1}{k})\\ \frac{dv_2}{dt}=D(v_1+v_3-2v_2)+q_1v_1-q_2v_2+v_2(1-\frac{u_2+v_2}{k})\\ \frac{dv_3}{dt}=D(v_2-v_3)+q_2v_2+v_3(1-\frac{u_3+v_3}{k})\\ u_i(0)=u_{i0},\quad v_i(0)=v_{i0},\quad i=1,2,3.\\ \end{array} \right. \end{aligned}$$
(2)

For Model II, it can also be shown that system (2) has a unique semi-trivial steady state of the form \((U^*,0)=(U_1^*,U_2^*,U_3^*,0,0,0)\), where \(U_i^*>0, i=1,2,3\).

Theorem 2

If \(q_1 \ge 1\) or \(\frac{q_2}{2}<q_1<1\), then \((U^*,0)\) is globally asymptotically stable for \(d>D\); i.e., the faster diffuser wins.

If \(q_1\ge 1\) and the diffusion rate of a species is small, then almost all of its individuals in patch 1 are washed out. Thus, small diffusion is not favored in this scenario. However, large diffusion will be selected as it can counterbalance the uni-dimensional drift by helping more individuals stay in patch 1. Similar intuition applies to the case \(q_2/2<q_1<1\), but the detail is more subtle: when diffusion rate is small, our analysis reveals that there will be more individuals in patch 2 than patch 1 when \(q_2/2<q_1<1\); i.e., the population in patch 1 is undermatching the resource more than in patch 2 (as carrying capacity in patches 1 and 2 is the same), so small diffusion is still not favored in this scenario.

Theorem 3

If \(0<q_1<1\) and \(q_2>2q_1\), there exists some \(d^*=d^*(q_1,q_2)>0\) which is an evolutionarily singular strategy. Moreover, this strategy is not a CvSS, and both zero and infinity dispersal rates are local CvSSs.

It is interesting that zero diffusion rate emerges as a local CvSS under the assumptions of Theorem 3. Suppose the diffusion is zero or very small. On one hand, when \(q_1<1\), the drift out of patch 1 is small enough to allow the population to persist in patch 1, which is a source. On the other hand, \(q_2>2q_1\) drive the population in patch 2 down and that in patch 3 up. Hence, patch 2 becomes a source and patch 3 becomes a sink. Moreover, diffusion takes individuals out of patch 1, but due to the uneven drift rates those individuals are more likely to end up in patch 3 (the sink) than in patch 2 (the source). Hence, increasing diffusion will move individuals from source patch to sink patch. Thus, small diffusion can be favored in this case as the drift forces more individuals to move from patch 1 to 2 to reduce the mismatch in patch 2.

When both zero and infinity dispersal rates are local CvSSs, a natural question is whether two competing populations can coexist. Our next result answers this question partially but positively:

Theorem 4

Fix any \(k, D, q_2 \ge 1\). Then there exists some \(\delta >0\) such that for any \(d\in (0, \delta )\), \(q_1\in (-d, \delta )\), Model II has a globally asymptotically stable positive steady state, denoted by \((U^\delta , V^\delta )\), which satisfies \((U^\delta ,V^\delta )\rightarrow ({\hat{U}}, {\hat{V}})\) as \(d\rightarrow 0\) and \(q_1\rightarrow 0\), where

$$\begin{aligned} {\hat{U}}:=(k-{\hat{V}}_2,0,0), \quad \text { and }\quad {\hat{V}}:=({\hat{V}}_2,{\hat{V}}_2,{\hat{V}}_3) \end{aligned}$$
(3)

such that \(({\hat{V}}_2, {\hat{V}}_3)\) is the unique positive solution of the two-patch system

$$\begin{aligned} \left\{ \begin{array}{l} D({\hat{V}}_3-{\hat{V}}_2)-q_2{\hat{V}}_2 +{\hat{V}}_2(1-{\frac{{\hat{V}}_2}{k}})=0 \\ D({\hat{V}}_2-{\hat{V}}_3)+q_2{\hat{V}}_2 +{\hat{V}}_3(1-{\frac{{\hat{V}}_3}{k}})=0. \end{array} \right. \end{aligned}$$
(4)

This result suggests that when the drift from patch 1 to patch 2 is very small, slow and fast diffusers can coexist in some interesting way: the slow diffuser will only occupy patch 1 and the fast diffuser is dominant in patches 2 and 3, but not in patch 1. Intuitively, the underlying mechanism for the coexistence is as follows: Consider the case \(d=0\) and \(q_1=0\) for the sake of clarity, so that patch 1 is disconnected from patches 2 and 3 for the species u. It turns out that, due to \(d=0\) and \(q_1=0\), the flux between patches 1 and 2 for the species v is also equal to zero. As a consequence, system (2) for patches 2 and 3 is reduced to a two-patch system for two competing species. It follows from previous work (Hamida 2017; Noble 2015) for two-patch models that the faster diffuser always out-competes the slower diffuser in patches 2 and 3, provided that \(q_2>0\). As patch 2 is at the upstream for the reduced two-patch model, the equilibrium distribution of species v, denoted by \(\hat{V_2}\), satisfies \(\hat{V_2}<k\); i.e., it undermatches the resource in patch 2. As there is no flux for species v between patches 1 and 2 and \(q_1=0\), the equilibrium distribution of species v at patch 1 is also equal to \(\hat{V_2}\), so that the equilibrium distribution of species u at patch 1 is given by \(k-\hat{V_2}>0\). The case of small \(d, q_1\) follows from a perturbation argument.

Note that \(q_1=0\) and small negative \(q_1\) are also covered by Theorem 4; the case of negative \(q_1\) applies to Model III.

2.3 Main Results of Model III

Model III assumes patch 1 is upstream, whereas patches 2 and 3 are downstream. Both species in patch 1 are transported to patches 2 and 3 by drift with rates \(q_1\) and \(q_2\), respectively. In this case we have the following ODE system for two competing species:

$$\begin{aligned} \left\{ \begin{array}{l} \frac{du_1}{dt}=d(u_2+u_3-2u_1)-(q_1+q_2)u_1+u_1(1-\frac{u_1+v_1}{k})\\ \frac{du_2}{dt}=d(u_1-u_2)+q_1u_1+u_2(1-\frac{u_2+v_2}{k})\\ \frac{du_3}{dt}=d(u_1-u_3)+q_2u_1+u_3(1-\frac{u_3+v_3}{k})\\ \frac{dv_1}{dt}=D(v_2+v_3-2v_1)-(q_1+q_2)v_1+v_1(1-\frac{u_1+v_1}{k})\\ \frac{dv_2}{dt}=D(v_1-v_2)+q_1v_1+v_2(1-\frac{u_2+v_2}{k})\\ \frac{dv_3}{dt}=D(v_1-v_3)+q_2v_1+v_3(1-\frac{u_3+v_3}{k})\\ u_i(0)=u_{i0},\quad v_i(0)=v_{i0},\quad i=1,2,3. \end{array} \right. \end{aligned}$$
(5)

The dynamics of (5) is more subtle than those for Models I and II. We first consider the global dynamics of Model III.

Theorem 5

If \(q_1=q_2>0\), then the semi-trivial steady state \((U^*,0)\) is globally asymptotically stable for \(d>D\).

Theorem 5 seems to agree with previous results for two-patch models that the faster diffuser always out-competes the slower diffuser (Hamida 2017; Noble 2015). The biological intuition is that both downstream patches are sinks under the assumption of Theorem 5; see also Corollary 6 (in “Appendix C”). Hence, any mutant in the upstream patch with smaller diffusion rate will more likely be pushed to two downstream sinks and thus cannot invade when rare.

Next we consider the local dynamics of Model III.

Theorem 6

If \(q_1,q_2>0\), \(| q_2-q_1 | \le \frac{1}{2}\) and \(\frac{1}{\sqrt{2}}\le \frac{q_2}{q_1}\le \sqrt{2}\), then the semi-trivial steady state \((U^*,0)\) is locally stable for \(d>D\) and unstable for \(d<D\).

Theorem 6 implies that infinite diffusion rate is a global CvSS when two drift rates are comparable. This is in the same spirit as Theorem 5 since both downstream patches are still sinks under the assumptions of Theorem 6; see also Corollary 8 (in “Appendix C”). In contrast, our next result shows that if two drift rates are not comparable, both zero and infinite diffusion rates are local CvSS.

Theorem 7

If \(1<q_1+q_2<(q_1-q_2)^2\), then there exists some \(d^*=d^*(q_1,q_2)>0\) which is an evolutionarily singular strategy. Moreover, \(d^*\) is not a CvSS, and both zero and infinity dispersal rates are local CvSSs.

To see why zero diffusion can be a local CvSS under the assumptions of Theorem 7, first fix \(q_1\) and choose \(q_2\) large. This will drain almost all individuals in patch 1 to drift to patch 3, so that patch 3 becomes a sink patch due to overcrowding. Subsequently, diffusion induces a net flux of individuals from patch 2 to patch 1, so that the population in patch 2 undermatches the resource. Hence, any mutant with smaller diffusion rate can invade when rare by exploiting patch 2, which is a source patch. The same intuitive reasoning applies to the general situation: for the range of \(q_i\) in Theorem 7, our numerical results suggest that one of the two downstream patches is a sink while the other becomes a source patch, and a mutant with smaller diffusion rate can invade when rare in the downstream source patch.

3 Conclusions

In this section, we first summarize the main analytical results, and then we provide a single framework to unify the main results for three models. The main findings are as follows, along with some predictions (see Sect. 4 for further discussions):

  • If all drift rates are identical, then the faster dispersal rate is selected across all three-patch models in which the drift network do not form a closed cycle. A conjecture is that this result holds for n-patch river networks with uniform carrying capacity and identical drift rates, provided that the drift network is not divergence-free (a drift network with identical drift rates is called divergence-free if each individual patch has the same number of upstream and downstream patches);

  • For general drift rates, infinite diffusion rate is a local CvSS for all three models. Biologically this makes good sense as with sufficiently fast dispersal, the spatial distribution of the species approaches the ideal free distribution. However, there are some notable differences for three models: For Model I, the faster dispersal is always favored and thus infinity is a global CvSS. For Models II and III, the answers depend upon the drifts rates: for some ranges of drift rates, infinity is a global CvSS (same as Model I), while for other ranges of drift rates, there exists some intermediate diffusion rate which is a singular strategy so that infinity is a local CvSS but not a global one. A conjecture is that the infinite diffusion rate is a local CvSS for n-patch river networks with uniform carrying capacity and general drift rates;

  • For Models II and III, when a singular strategy (that is neither zero nor infinity) exists, it is not a local CvSS (Numerical simulation suggests that it is not an ESS either). In fact, bi-stability phenomenon happens, i.e., both infinity and zero diffusion rates are local CvSSs;

  • For Models II and III, when bi-stability occurs, it is possible for two competing populations with different dispersal rates to coexist, by varying the drift rates between patches. A conjecture is that any coexistence steady state for Models II–III, if exists, is globally stable.

Next, we provide a single framework to unify the main results for Models I–III. Our idea is to use a single system of ODEs to describe all three models. Without loss of generality, consider system (2) in the \(q_1-q_2\) plane, allowing the drift rates in system (2) to take both positive and negative values. That is, we divide the \(q_1-q_2\) plane into 4 quadrants. Then the first quadrant of Fig. 2 corresponds to Model II with nonnegative drift rates.

  • First quadrant: Theorem 2 implies that the faster diffuser always wins for \(q_1, q_2\) in the red region; for the blue region, Theorem 3 ensures the existence of an evolutionarily singular strategy, where both zero and infinity dispersal rates are local CvSSs.

  • Second quadrant: With \(q_1<0\) and \(q_2>0\) in system (2), the directed flows are from patch 2 to patches 1 and 3. Hence, this corresponds to Model III with patches 1 and 2 switched. Theorem 6 implies that the faster diffuser wins for \(q_1, q_2\) in the red region; for the blue region, Theorem 7 ensures the existence of an evolutionarily singular strategy, in which both zero and infinity dispersal rates are local CvSSs.

  • Third quadrant: With \(q_1<0\) and \(q_2<0\) in system (2), the directed flows are from patch 3 to 2, and from patch 2 to 1. Hence, this corresponds to Model II with patches 1 and 3 switched. Hence the red and blue regions are symmetric to those in the 1st quadrant with respect to the line \(q_1+q_2=0\).

  • Fourth quadrant: With \(q_1>0\) and \(q_2<0\) in system (2), the directed flows are from patches 1 and 3 to patch 2. Hence, this corresponds to Model I with patches 2 and 3 switched. Theorem 1 implies that the faster diffuser always wins for \(q_1, q_2\) in Fourth quadrant.

These discussions suggest that the \(q_1-q_2\) plane can also be divided into three colored regions as in Fig. 2: For the red region, the infinite diffusion rate is a global CvSS; in the blue region, both zero and infinite diffusion rates are local CvSSs; for the white region, numerical simulations suggest that the infinite diffusion rate is a local CvSS but might not be a global one, and the zero diffusion rate might not even be a local CvSS.

Fig. 2
figure 2

The dynamics for Models I–III. The red colored regions correspond to the ranges of \(q_1, q_2\) for which the infinite diffusion rate is a global CvSS; in the blue colored regions, there is at least one evolutionarily singular strategy and both infinity and zero diffusion rates are local CvSSs; the dynamics of Model III in the white colored regions is not fully determined theoretically (Color figure online)

4 Discussion and Numerical Results for Four-patch Models

In this section, we will discuss possible extensions to n-patch river network models and raise some conjectures on the evolution of faster dispersal. We will also address the issue of the invasion of slowly diffusing populations and propose to study the coexistence of slow and fast diffusing competing populations.

4.1 Evolution of Fast Dispersal in n-patch Model

Theorems 1,  2 and 5 show that if \(q_1=q_2\), the faster diffusing population always wins the competition for Models I–III. In particular, infinity as a diffusion rate is a global CvSS for Model I and also for wider ranges of parameters in both Models II and III; see Theorems 2 and 6.

Consider the general n-patch river model, i.e.,

$$\begin{aligned} {\left\{ \begin{array}{ll} \frac{du_i}{dt} =d\sum _{j=1}^n m_{ij} u_j +\sum _{j=1}^n q_{ji}u_j +u_i(1-\frac{u_i+ v_i}{k_i}), \quad 1\le i\le n,\\ \frac{dv_i}{dt} =D\sum _{j=1}^n m_{ij} v_j +\sum _{j=1}^n q_{ji}v_j +v_i(1-\frac{u_i+ v_i}{k_i}), \quad 1\le i\le n.\\ \end{array}\right. } \end{aligned}$$

Here the connectivity matrix \(M:=(m_{ij})\) is assumed to be symmetric, \(m_{ij}=m_{ji}=1\) when two patches i and j are directly connected, \(m_{ij}=m_{ji}=0\) when they are not directly connected, and \(m_{ii}=-\sum _{j\not =i} m_{ij}\). The drift matrix \(Q:=(q_{ij})\) satisfies \(q_{ii}=-\sum _{j\not =i} q_{ji}\), \(q_{ij}>0\) when patches i and j are connected and the directed flow is from patch i to j, and \(q_{ij}=0\) otherwise. The case \(m_{ij}=1\) but \(q_{ij}=0\) refers to the scenario when patches ij are directly connected but there is no directed passive flow in between. Note that under our assumptions, the dominant eigenvalue of Q is zero, with left eigenvector being \((1,\ldots ,1)\), i.e.,

$$\begin{aligned} (1,\ldots ,1)Q=0. \end{aligned}$$

The positive constant \(k_i\) is the carrying capacity of patch i.

Definition 1

We say that the drift matrix \(Q=(q_{ij})\) is divergence-free if its right eigenvector, corresponding to the zero eigenvalue, is given by \((1,\ldots ,1)^T\), i.e.,

$$\begin{aligned} \sum _{j: j \ne i} q_{ij} = \sum _{j: j\ne i} q_{ji} \quad \text { holds for each }i. \end{aligned}$$

Conjecture 1

If all positive drift rates are equal, the carrying capacity is the same for all patches, and the drift network \((q_{ij})\) is not divergence-free, then the faster disperser always wins.

For general drift rates, Theorem 1 shows that infinite dispersal rate is a global CvSS for Model I, while for Models II and III, Lemmas 15 (in “Appendix B”) and 36 (in “Appendix C”) find that infinity is always a local CvSS.

Conjecture 2

For n-patch model with general drift rates, when the drift matrix is not divergence-free and that the carrying capacity is the same for all patches, the infinite diffusion rate is always a local CvSS.

From the biological point of view, when \(k_i = k\) for all i, for a single species with sufficiently fast diffusion, its equilibrium will be close to \((k,\ldots , k)\), which is an ideal free distribution. Heuristically, if a strategy can help organisms reach the ideal free distribution at equilibrium, then the strategy is likely to be a local ESS and/or CvSS; see (Cantrell et al. 2007, 2012, 2017; Lou 2019). Again, we may need to exclude the exceptional case \((1,\ldots ,1)^T\) being a right eigenvector of matrix Q.

To support the above predictions for n-patch models, we performed some numerical simulations for the following four-patch models with the network topology as shown in Fig. 3:

Fig. 3
figure 3

The two-way blue arrows represent the dispersal of species between connected patches, the one-way red arrows represent the uni-directional drift. The parameter dD are dispersal rates for two competing species, and the parameters \(q, {\tilde{q}}\) are drift rates. Patches 1–3 form a loop. Patch 4 is at the upstream in (a) and it is the downstream patch in (c). There is no drift between patches 3 and 4 in (b) (Color figure online)

For 4-patch model with topology Fig. 3a, our simulation results suggest that for any \({\tilde{q}}>0\), the faster diffusing species always wins the competition, and the conclusion is independent of drift rates. In particular, the faster dispersal rate is selected when \({\tilde{q}}=q\), which is consistent with Conjectures 1 and 2.

Figure 3b can also be viewed as Fig. 3a, c with \({\tilde{q}}=0\). For this special case, \((k,\ldots , k)\) is the unique positive equilibrium for the corresponding single species model. This gives an example of the exceptional case discussed earlier for n-patch models. As predicted earlier, our numerical simulations show that two populations with different dispersal rates coexist, i.e., the faster diffusing species does not win the competition in this exceptional case.

The PIP for 4-patch model with topology Fig. 3c is shown in Fig. 4. We take \(d\in [0,2]\) and \(D\in [0,2]\), and then we discretize the interval [0, 2] with the uniform step \(\varDelta =0.02\). The parameter values \((k_1,k_2,k_3,k_4)\) are set to be (7, 7, 7, 7) and \(q=1\), \({\tilde{q}}\) ranges from 0.01 to 2000. Our simulations (see Fig. 4) suggest more complicated dynamics: when \({\tilde{q}}\in (0, q]\), the faster diffusing population still wins. In particular, the fast dispersal rate is selected when \({\tilde{q}}=q\), which is consistent with Conjectures 1 and 2. However, for \({\tilde{q}}\) larger than q, there are two evolutionarily singular strategies, one is a local ESS and CvSS, and the other is neither an ESS not CvSS. Furthermore, the infinite diffusion rate remains as a local CvSS as predicted, while the zero diffusion rate is not a local CvSS.

Fig. 4
figure 4

Pairwise invasion plots (PIPs) for the four-patch model with network topology Fig. 3c. \(k_1=k_2=k_3=k_4=7\), \(q=1\) and \({\tilde{q}}\) ranges from 0.01 to 2000. The horizontal axis is d and the vertical axis is D. The black regions represent the range of (dD) for which \((U^* ,0)\) is locally stable

4.2 Evolution of Slow Dispersal and Coexistence

Theorems 3 and 7 illustrate the existence of evolutionarily singular strategy for Models II and III, respectively. These singular strategies are not local CvSSs, and numerical simulations suggest they are not local ESSs either. In fact, Lemmas 16 (in “Appendix B”) and 37 (in “Appendix C”) show that zero diffusion rate can be a local CvSS for some parameter ranges in both Models II and III.

A natural question for general n-patch models is when zero diffusion rate can be a local CvSS. The analysis of Model III reveals that if there are more than one downstream patches, then it is possible for one of them to be a source patch, so that a mutant with slow diffusion rate can invade when rare in this source patch. For n-patch models it will be of interest to find sufficient conditions on the existence of some downstream source patch, by taking into account of the river network topology, so that slow diffusing populations can invade such source patch when rare.

For general n-patch models, when both zero and infinite diffusion rate are local CvSSs, it is natural to inquire whether slow and fast diffusers can coexist. It will be of interest to generalize Theorem 4 to n-patch models and to reveal the impact of network topology on the coexistence of competing species.