Skip to main content
Log in

Dynamics of Phytoplankton Communities Under Photoinhibition

  • Original Article
  • Published:
Bulletin of Mathematical Biology Aims and scope Submit manuscript

Abstract

We analyzed a model of phytoplankton competition for light in a well-mixed water column. The model, proposed by Gerla et al. (Oikos 120:519–527, 2011), assumed inhibition of photosynthesis at high irradiance (photoinhibition). We described the global behavior through mathematical analyses, providing a general solution to the multi-species competition for light with photoinhibition. We classified outcomes of 2- and 3-species competitions as examples, and evaluated feasibility of the theoretical predictions using empirical relationships between photosynthetic production and irradiance. Numerical simulations with published pI curves indicate that photoinhibition may often lead to strong Allee effects and competitive facilitation among species. Hence, our results suggest that photoinhibition may play a major role in organizing phytoplankton communities.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  • Alderkamp, A. C., de Baar, H. J. W., Visser, R. J. W., & Arrigo, K. R. (2010). Can photoinhibition control phytoplankton abundance in deeply mixed water columns of the Southern Ocean? Limnol. Oceanogr., 55, 1248–1264.

    Article  Google Scholar 

  • Andrews, J. F. (1968). A mathematical model for the continuous culture of microorganisms utilizing inhibitory substrates. Biotechnol. Bioeng., 10, 707–723.

    Article  Google Scholar 

  • Baastrup-Spohr, L., & Staehr, P. A. (2009). Surface microlayers on temperate lowland lakes. Hydrobiologia, 625, 43–59.

    Article  Google Scholar 

  • Basterretxea, G., & Aristegui, J. (2000). Mesoscale variability in phytoplankton biomass distribution and photosynthetic parameters in the Canary–NW African coastal transition zone. Mar. Ecol. Prog. Ser., 197, 27–40.

    Article  Google Scholar 

  • Bischof, K., Hanelt, D., & Wiencke, C. (1998). UV-radiation can affect depth-zonation of Antarctic macroalgae. Mar. Biol., 131, 597–605.

    Article  Google Scholar 

  • Bouchard, J. N., Roy, S., & Campbell, D. A. (2006). UVB effects on the photosystem II-D1 protein of phytoplankton and natural phytoplankton communities. Photochem. Photobiol., 82, 936–951.

    Article  Google Scholar 

  • Butler, G. J., & Wolkowicz, G. S. K. (1985). A mathematical model of the chemostat with a general class of functions describing nutrient uptake. SIAM J. Appl. Math., 45, 138–151.

    Article  MathSciNet  MATH  Google Scholar 

  • Eilers, P. H. C., & Peeters, J. C. H. (1988). A model for the relationship between light intensity and the rate of photosynthesis in phytoplankton. Ecol. Model., 42, 199–215.

    Article  Google Scholar 

  • Elser, J. J., & Kimmel, B. L. (1985). Photoinhibition of temperate lake phytoplankton by near-surface irradiance: evidence from vertical profiles and field experiments. Eur. J. Phycol., 21, 419–427.

    Google Scholar 

  • Gerla, D. J., Mooij, W. M., & Huisman, J. (2011). Photoinhibition and the assembly of light-limited phytoplankton communities. Oikos, 120, 359–368.

    Article  Google Scholar 

  • Han, B. P. (2002). A mechanistic model of algal photoinhibition induced by photodamage to photosystem-II. J. Theor. Biol., 214, 519–527.

    Article  Google Scholar 

  • Henley, W. J. (1993). Measurement and interpretation of photosynthetic light-response curves in algae in the context of photoinhibition and diel changes. Eur. J. Phycol., 29, 729–739.

    Article  Google Scholar 

  • Huisman, J. (1997). The struggle for light. Ph.D. thesis, Rijksuniversiteit Groningen.

  • Huisman, J., & Weissing, F. J. (1994). Light-limited growth and competition for light in well-mixed aquatic environments: an elementary model. Ecology, 75, 507–520.

    Article  Google Scholar 

  • Huisman, J., & Weissing, F. J. (1995). Competition for nutrients and light in a mixed water column: a theoretical analysis. Am. Nat., 146, 536–564.

    Article  Google Scholar 

  • Klausmeier, C. A., & Litchman, E. (2001). Algal games: the vertical distribution of phytoplankton in poorly mixed water columns. Limnol. Oceanogr., 46, 1998–2007.

    Article  Google Scholar 

  • Kohyama, T. (1993). Size-structured tree populations in gap-dynamic forest: the forest architecture hypothesis for the stable coexistence of species. J. Ecol., 81, 131–143.

    Article  Google Scholar 

  • Litchman, E., & Klausmeier, C. A. (2008). Trait-based community ecology of phytoplankton. Annu. Rev. Ecol. Evol. Syst., 39, 615–639.

    Article  Google Scholar 

  • Litchman, E., & Neale, P. J. (2005). UV effects on photosynthesis, growth and acclimation of an estuarine diatom and cryptomonad. Mar. Ecol. Prog. Ser., 300, 53–62.

    Article  Google Scholar 

  • Litchman, E., Neale, P. J., & Banaszak, A. T. (2002). Increased sensitivity to ultraviolet radiation in nitrogen-limited dinoflagellates: photoprotection and repair. Limnol. Oceanogr., 47, 86–94.

    Article  Google Scholar 

  • McGill, B. J., Enquist, B. J., Weiher, E., & Westoby, M. (2006). Rebuilding community ecology from functional traits. Trends Ecol. Evol., 21, 178–185.

    Article  Google Scholar 

  • Mellard, J. P., Yoshiyama, K., Litchman, E., & Klausmeier, C. A. (2011). The vertical distribution of phytoplankton in stratified water columns. J. Theor. Biol., 269, 16–30.

    Article  MathSciNet  Google Scholar 

  • Muller, E. B. (2010). Synthesizing units as modeling tool for photosynthesizing organisms with photoinhibition and nutrient limitation. Ecol. Model., 222, 637–644.

    Article  Google Scholar 

  • Oliver, R. L., Whittington, J., Lorenz, Z., & Webster, I. T. (2003). The influence of vertical mixing on the photoinhibition of variable chlorophyll a fluorescence and its inclusion in a model of phytoplankton photosynthesis. J. Plankton Res., 25, 1107–1129.

    Article  Google Scholar 

  • Ryabov, A. B., Rudolf, L., & Blasius, B. (2010). Vertical distribution and composition of phytoplankton under the influence of an upper mixed layer. J. Theor. Biol., 263, 120–133.

    Article  MathSciNet  Google Scholar 

  • Schwaderer, A. S., Yoshiyama, K., de Tezanos Pinto, P., Swenson, N. G., Klausmeier, C. A., & Litchman, E. (2011). Eco-evolutionary differences in light utilization traits and distributions of freshwater phytoplankton. Limnol. Oceanogr., 56, 589–598.

    Article  Google Scholar 

  • Stomp, M., Huisman, J., Voros, L., Pick, F. R., Laamanen, M., Haverkamp, T., & Stal, L. J. (2007). Colourful coexistence of red and green picocyanobacteria in lakes and seas. Ecol. Lett., 10, 290–298.

    Article  Google Scholar 

  • Tyystjärvi, E. (2008). Photoinhibition of photosystem II and photodamage of the oxygen evolving manganese cluster. Coord. Chem. Rev., 252, 361–376.

    Article  Google Scholar 

  • Weissing, F. J., & Huisman, J. (1994). Growth and competition in a light gradient. J. Theor. Biol., 168, 323–336.

    Article  Google Scholar 

  • Yoshiyama, K., Mellard, J. P., Litchman, E., & Klausmeier, C. A. (2009). Phytoplankton competition for nutrients and light in a stratified water column. Am. Nat., 174, 190–203.

    Article  Google Scholar 

  • Yoshiyama, K., & Nakajima, H. (2002). Catastrophic transition in vertical distributions of phytoplankton: alternative equilibria in a water column. J. Theor. Biol., 216, 397–408.

    Article  MathSciNet  Google Scholar 

  • Zonneveld, C. (1998). Photoinhibition as affected by photoacclimation in phytoplankton: a model approach. J. Theor. Biol., 193, 115–123.

    Article  Google Scholar 

Download references

Acknowledgements

We are grateful to Jef Huisman and an anonymous referee for their helpful comments and suggestions which led to an improvement of our original manuscript. Kohei Yoshiyama would like to thank the National Center for Theoretical Science, National Tsing-Hua University, Taiwan, for its financial support and kind hospitality during his visit there.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Sze-Bi Hsu.

Appendix

Appendix

1.1 A.1 Basic Properties

We will demonstrate the shape of function

where p(I) satisfies p(0)=0, p(I)∈(0,∞) for 0<II in, dp/dI<∞, dp/dI>0 for 0≤I<I opt, and dp/dI<0 for I opt<II in. We claim that g(0)=0 and g(I in)=p(I in), and there is a unique point \(\hat{x}\) such that \(d g(\hat{x})/dx = 0\), and dg/dx>0 for \(x\in[0,\hat{x})\); dg/dx<0 for \(x \in(\hat{x}, I_{\mathrm{in}}]\).

The proof of g(0)=0 is straightforward:

because \(\int_{0}^{I_{\mathrm{in}}} \frac{p(I)}{I}\,dI\) is bounded and lim x→0(lnI in−lnx)=∞. Likewise, g(I in)=p(I in) can be proved by

Next, we prove that g(x) is a unimodal function with a peak at \(\hat{x} \in(0, I_{\mathrm{opt}})\). The derivative of g(x) is

Let

Because p(0)=0 from the assumption, we have

Therefore, g 1(0)<g 2(0) and g 1(I in)=g 2(I in)=0.

The derivatives of g 1(x) and g 2(x) are:

(12)
(13)

Because p′(x)<0 for x∈(I opt,I in], we have \(|g_{1}^{\prime}(I_{\mathrm{in}}^{-})|>|g_{2}^{\prime}(I_{\mathrm{in}}^{-})|\). Since g 1(I in)=g 2(I in), it follows that g 1(x)>g 2(x) for x near I in. By the intermediate value theorem, there exists a point \(\hat{x} \in(0,I_{\mathrm{in}})\) such that \(g_{1}(\hat{x})=g_{2}(\hat{x})\).

From \(g_{1}(\hat{x})=g_{2}(\hat{x})\), g 1(I in)=g 2(I in), there exists a point \(\bar{x} \in(\hat{x},I_{\mathrm{in}})\) such that \(g_{1}^{\prime}(\bar{x})=g_{2}^{\prime}(\bar{x})\) according to Rolle’s theorem. From (12) and (13), we have \(p^{\prime}(\bar{x})=0\). From assumption of p(x), it follows that \(\bar{x}=I_{\mathrm{opt}}\). Hence, \(\hat{x}<I_{\mathrm{opt}}<I_{\mathrm{in}}\).

Suppose there exists another point \(\hat{x}^{\prime}\in(0,I_{\mathrm{in}})\) with \(g_{1}(\hat{x}^{\prime})=g_{2}(\hat{x}^{\prime})\). If \(\hat{x}^{\prime}\in(\hat{x},I_{\mathrm{opt}}]\), then by Rolle’s theorem there exists a point \(s\in(\hat{x},\hat{x}^{\prime})\) such that p′(s)=0. It contradicts to that p′(I)>0 for 0≤I<I opt. If \(\hat{x}^{\prime}\in(I_{\mathrm{opt}},I_{\mathrm{in}})\), then by Rolle’s theorem there exists a point s∈(I opt,I in) such that p′(s)=0. It contradict to that p′(I)<0 for 0≤I opt<I<I in. Thus, \(\hat{x}\) is the only point with \(g_{1}(\hat{x})=g_{2}(\hat{x})\).

Next, we prove the following theorem. In the following, we let L(t)=I out(t) for convenience. From (11),

(14)

Theorem 2

The solutions of (10) are positive and bounded.

Proof

Assume there exists t 1>0 and some i∈{1,2,…,n} such that

By reversing time, let τ=−t and we consider backward behavior of the solution of (10) with initial data x i (0)=0,x j (0)=x j (t 1)>0. It follows that x i (τ)=0 for all τ<0. By the uniqueness of ordinary differential equations, we have x i (−t 1)=0, a contradiction. Thus, the solutions are positive if the initial condition is in Ω.

To prove the boundedness of solution, we consider the differential inequalities

where

with \(g_{i}(L_{\max,i})=\max_{s\in[0,I_{0}]}g_{i}(s)\). Let y(t) be the solution of

$$y'=\bigl[G_i\bigl(I_0 \exp(-k_iy)\bigr)-d_i\bigr] y. $$

Then lim t→∞ y(t)=y i , where y i satisfies G i (I 0exp(−k i y i ))−d i =0. Hence, for given small ϵ, x i (t)≤y i +ϵ for all t large. Hence, x i (t) is bounded for all time t for i=1,2,…,n. □

The next theorem says that if the light intensity is weak enough, then some species die out.

Theorem 3

If λ i >I 0, then lim t→∞ x i (t)=0.

Proof

Since L(t)≤I 0 for all t≥0, we have L(t)<λ i for all t≥0. From g i (L(t))≤G i (L(t)),

$$\frac{x_i'}{x_i}=g_i\bigl(L(t)\bigr)-d_i\leq G_i\bigl(L(t)\bigr)-d_i\leq G_i(I_0)-d_i<0, $$

then x i (t)→0 as t→∞. □

1.2 A.2 Local Stability of Equilibria

For simplicity, let

$$f_i\bigl(x_1(t),x_2(t), \ldots,x_n(t)\bigr):=\bigl[g_i\bigl(L(t) \bigr)-d_i\bigr]x_i(t). $$

We denote the Jacobian of (10) at an equilibrium E is \(J(E)=[m_{ij}]\in\mathbb{R}^{n\times n}\), where

For the equilibrium E 0, the Jacobian at E 0 is

$$J(E_0) = \left [ \begin{array}{c@{\quad}c@{\quad}c@{\quad}c@{\quad}c} g_1(I_0)-d_1 & 0 & 0 & \ldots&0 \\ 0 & g_2(I_0)-d_2 &0 &\ldots&0 \\ \vdots\\ 0 & 0 &\ldots & 0 & g_n(I_0)-d_n \\ \end{array} \right ]. $$

Obviously the eigenvalues of J(E 0) are g i (I 0)−d i , for i=1,2,…,n.

If I 0S, then g i (I 0)−d i <0 for all i and E 0 is locally asymptotically stable.

If I 0S, then there exists some i such that I 0∈(λ i ,μ i ). Hence, g i (I 0)−d i >0, and E 0 is unstable. If there exists some j such that I 0∉(λ j ,μ j ), then J(E 0) has negative eigenvalues and E 0 is saddle with the local stable manifold

$$W_\mathrm{loc}^s(E_0)= \Biggl\{\sum _{i=1}^n c_ie_i: c_i=0 \mbox{ except some }j \mbox{ with } I_0\notin( \lambda_j,\mu_j). \Biggr\}, $$

where e i is the eigenvector corresponding to the eigenvalue of (g i (I 0)−d i ). Hence, the dimension of W s(E 0), the stable manifold of E 0, is at most (n−1) and

$$W^s(E_0)\subset\bigl\{(x_1, \ldots,x_n): x_i=0 \mbox{ if } I_0\in( \lambda_{i},\mu_{i})\bigr\}. $$

Therefore, W s(E 0)∩Ω=∅.

For the equilibria \(E_{\lambda_{r}}\), the Jacobian evaluated at \(E_{\lambda_{r}}\) is

$$J(E_{\lambda_r}) = \left [ \begin{array}{c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad }c@{\quad}c@{\quad}c} m_{11} & 0 &\ldots&\ldots& \ldots&\ldots&\ldots& \ldots&0 \\ 0 & m_{22} &0 &\ldots&\ldots& \ldots&\ldots&\ldots&0 \\ \vdots\\ 0 &\ldots &0 &m_{r-1,r-1} &0 &\ldots&\ldots&\ldots &0\\ m_{r1} &\ldots&\ldots &m_{r,r-1} &m_{r,r} &m_{r,r+1} &\ldots&\ldots &m_{rn}\\ 0 &\ldots&\ldots&\ldots&0 &m_{r+1,r+1} &0 &\ldots&0\\ \vdots\\ 0 &\ldots&\ldots&\ldots&\ldots&\ldots&\ldots& 0 & m_{nn} \\ \end{array} \right ], $$

where

The eigenvalues of \(J(E_{\lambda_{r}})\) are \(-k_{r}x_{\lambda_{r}}g_{r}'(\lambda_{r})\lambda_{r}<0\), and g j (λ r )−d j , for all jr.

When λ r is the endpoint of a component of S, we have that g j (λ r )−d j <0 for jr. Hence, \(E_{\lambda_{r}}\) is locally asymptotically stable.

If λ r is not the endpoint of a component of S, then there exists j such that λ r ∈(λ j ,μ j ). The eigenvector corresponding to the negative eigenvalue \(-k_{r}x_{\lambda_{r}}g_{r}'(\lambda_{r})\lambda_{r}\) is

$$v_r=(0,\ldots,0,x_r,0,\ldots,0),\quad x_r>0 \mbox{ in the }r\mbox{th component}, $$

and the eigenvectors corresponding to the negative eigenvalues g i (λ r )−d i are

$$v_i=(0,\ldots,0,k_r,0,\ldots,0,-k_i,0, \ldots,0), $$

where k r , k j >0 are in the ith , and rth component of vector v i , respectively. We know that \(E_{\lambda_{r}}\) is saddle and the local stable manifold is

$$W_\mathrm{loc}^s(E_{\lambda_r})= \Biggl\{E_{\lambda_r}+\sum _{i=1}^nc_i v_i:\ c_i=0 \mbox{ except } c_r \mbox{ and some } j\mbox{ with } \lambda_r\notin(\lambda_j,\mu_j) \Biggr\}. $$

Hence, the stable manifold of \(E_{\lambda_{r}}\) is

$$W^s(E_{\lambda_r})\subset\bigl\{(x_1, \ldots,x_n): x_i=0 \mbox{ if } \lambda_r\in( \lambda_{i},\mu_{i})\bigr\}, $$

and \(W^{s}(E_{\lambda_{r}})\cap\varOmega=\emptyset\).

For the equilibria \(E_{\mu_{r}}\), the Jacobian evaluated at \(E_{\mu_{r}}\) is \(J(E_{\mu_{r}})\), the structure is similar to \(J(E_{\lambda_{r}})\) and

The eigenvalues are \(-k_{r}x_{\mu_{r}}g_{r}'(\mu_{r})\mu_{r}>0\), an g j (μ r )−d j , for all jr.

For the case μ r is the endpoint of a component of S, then g j (μ r )−d j <0 for all jr. Hence, \(E_{\mu_{r}}\) is saddle with one dimensional unstable manifold

$$W^u(E_{\mu_r})= \bigl\{E_{\mu_r}+sv_r: v_r=(0,\ldots,0,x_r,0,\ldots,0), x_r>0 \mbox{ in the }r\mbox{th component} \bigr\}. $$

The eigenvectors corresponding to the negative eigenvalues g j (λ r )−d j for all jr are

$$v_j=(0,\ldots,0,k_r,0,\ldots,0,-k_j,0, \ldots,0), $$

where k r , k j >0 are in the jth, and rth components of the vector v j , respectively. \(E_{\mu_{r}}\) is saddle which stable manifold \(W^{s}(E_{\mu_{r}})\) is tangent to

$$W_\mathrm{loc}^s(E_{\mu_r})= \Biggl\{E_{\mu_r}+\sum _{j=1}^{n}c_jv_j: c_r=0, c_j\neq0, \mbox{ for }j\neq r \Biggr\}, $$

at \(E_{\mu_{r}}\). Thus, the (n−1)-dimensional stable manifold of \(E_{\mu_{r}}\) satisfies

$$W^s(E_{\mu_r})\subset\bigl\{(x_1,x_2, \ldots,x_n), x_i>0 \mbox{ for all } i\bigr\}, $$

and \(W^{s}(E_{\mu_{r}})\cap\varOmega\neq\emptyset\).

For the case μ r is not the endpoint of a component of S, then there exists j such that μ r ∈(λ j ,μ j ) and g j (λ r )−d j >0. If there exists some i such that μ r ∈(λ i ,μ i ), then \(E_{\mu_{r}}\) is saddle which stable manifold \(W^{s}(E_{\mu_{r}})\) is tangent to

$$W_\mathrm{loc}^s(E_{\mu_r})= \Biggl\{E_{\mu_r}+\sum _{j=1}^nc_jv_j: c_j=0 \mbox{ except } c_i \mbox{ with } \mu_r\notin(\lambda_i,\mu_i) \Biggr\}, $$

at \(E_{\mu_{r}}\). Hence, the stable manifold of \(E_{\mu_{r}}\) is

$$W^s(E_{\mu_r})\subset\bigl\{(x_1, \ldots,x_n): x_i=0 \mbox{ if } \mu_r\in( \lambda_{i},\mu_{i})\bigr\}, $$

and \(W^{s}(E_{\mu_{r}})\cap\varOmega=\emptyset\).

1.3 A.3 The Proof of Theorem 1

To prove Theorem 1, we need the following three lemmas. We note that the following proofs are similar to those in Butler and Wolkowicz (1985). We present them for the sake of completeness of the paper.

Lemma 4

If lim t→∞ x i (t)>0, then lim t→∞ L(t)=λ i or μ i and lim t→∞ x j (t)=0 for ji.

Proof

Since lim t→∞ x i (t) exists and it is positive, and \(|x_{i}''(t)|\) is bounded, then \(x_{i}'(t)\) converges to 0 as t goes to infinity, i.e.,

$$g_i\bigl(L(t)\bigr)-d_i\to0 \quad\mbox{as } t\to \infty. $$

Therefore, lim t→∞ L(t)=λ i or μ i .

For ji, we prove lim t→∞ x j (t)=0 by contradiction. First, we assume lim t→∞ x j (t)>0, then by the similar argument as above, we obtain that lim t→∞ L(t)=λ j or μ j , a contradiction. Thus, lim t→∞ x j (t) does not exist and lim sup t→∞ x j (t)>0. Then there exists a subsequence t m increases to infinity as m goes to infinity, and x j (t m ) converges to lim sup t→∞ x j (t) and \(x_{j}'(t_{m})=0\). Hence, g j (L(t m ))−d j =0 and L(t m )=λ j or μ j , a contradiction to lim t→∞ L(t)=λ i or μ i . Thus, we have lim t→∞ x j (t)=0. □

Lemma 5

If lim t→∞ L(t)=γ, then γ∈{I 0, the endpoints of a component of S}.

  1. 1.

    If γ=I 0, then I 0S, and lim t→∞ x i (t)=0 for all i.

  2. 2.

    If γ=λ i or μ i , the endpoint of a component of S, then γ<I 0 and \(\lim_{t\to\infty}x_{i}(t)=x_{\lambda_{i}}\) or \(x_{\mu_{i}}\) and lim t→∞ x j (t)=0 for ji.

Proof

We prove by contradiction. If not, then γ∉{I 0, the endpoints of a component of S}. There are two possibilities: γS and γS.

If γS, from the assumption lim t→∞ L(t)=γ, then for ϵ>0 there exists some i and T ϵ >0 such that L(t)⊂(λ i ,μ i ) for tT ϵ . It follows that \(x_{i}'(t)\geq0\) for tT ϵ , and from the fact x i (t) is bounded above, then lim t→∞ x i (t)>0. By Lemma 4, we have that γ=lim t→∞ L(t)=λ i or μ i . Since γ is not endpoints of S, there exists ji such that γ∈(λ j ,μ j ). It follows that L(t)⊂(λ j ,μ j ) for all large t. By similar argument as above, we have lim t→∞ L(t)=λ j or μ j , a contradiction.

If γS, then for ϵ>0 there exists T ϵ >0 s.t. L(t)∈(γϵ,γ+ϵ)⊂S c for tT ϵ . Hence \(\frac{x_{i}'(t)}{x_{i}(t)}=g_{i}(L(t))-d_{i}<0\) for all i for tT ϵ . Therefore, lim t→∞ x i (t)=0 for all i, and lim t→∞ L(t)=I 0, a contradiction.

  1. 1.

    Let γ=I 0. Assume I 0S and from the convergence of L(t), there exists i such that L(t)∈(λ i ,μ i ) for all t is large. By similar argument as above, we have lim t→∞ x i (t)>0 and lim t→∞ L(t)=λ i or μ i , it is a contradiction. Thus, I 0S.

    Now we prove lim t→∞ x i (t)=0 for all i by contradiction. First we assume that there exists i such that lim t→∞ x i (t)>0. Then lim t→∞ L(t)=λ i or μ i , a contradiction. Assume lim t→∞ x i (t) does not exist and lim sup t→∞ x i (t)>0. Then there exists a sequence {t m } increases to infinity as t goes to infinity such that \(x_{i}'(t_{m})=0\) and lim m→∞ x i (t m )=lim sup t→∞ x i (t). It follows that g i (L(t m ))−d i =0 and L(t m )=λ i or μ i for all m, a contradiction to γ=I 0. Thus, lim t→∞ x i (t)=0 for all i.

  2. 2.

    It is clear that γ<I 0, since L(t)≤I 0 for all t. lim t→∞ x j (t)=0 for ji follows from the above argument. Thus, it follows that \(\lim_{t\to\infty}L(t)= \lim_{t\to\infty}[I_{0} e^{-k_{i} x_{i}(t)}]=\lambda_{i}\) or μ i , or equivalently \(\lim_{t\to\infty}x_{i}(t)=x_{\lambda_{i}}\) or \(x_{\mu_{i}}\). □

Lemma 6

L(t) converges as t goes to infinity.

Proof

If not, then there exist increasing sequences {t m },{τ m } such that

$$L'(t_m)=0, \lim_{m\to\infty}L(t_m)= \limsup_{t\to\infty}L(t):=\overline{L}, $$
$$L'(\tau_m)=0, \lim_{m\to\infty}L( \tau_m)=\liminf_{t\to\infty }L(t):=\underline{L}. $$

Note that there are some i∈{1,2,…,n} such that x i (t) do not tend to zero. Since

$$L'(t_m)=-L(t_m) \Biggl[\sum _{i=1}^{n}k_ix_i'(t_m) \Biggr]=0, $$

for each m there are some j m ∈{1,2,…,n} satisfies \(x_{j_{m}}'(t_{m})\geq0\). There exists some j such that j m =j for infinitely many m. For this j, we choose a subsequence of {t m }, also named {t m }, such that \(x_{j}'(t_{m})\geq0\). It follows that L(t m )∈[λ j ,μ j ] for all m and \(\overline{L}\in[\lambda_{j},\mu_{j}]\). Similarly, we can find some k and a subsequence of {τ m }, also named {τ m }, such that L(τ m )∈[λ k ,μ k ] for all m and \(\underline{L}\in[\lambda_{k},\mu_{k}]\).

If \(\overline{L}\in[\lambda_{j},\mu_{j}]\subset[\lambda_{p_{1}},\mu_{q_{1}}]\) and \(\underline{L}\in[\lambda_{k},\mu_{k}]\subset[\lambda_{p_{2}},\mu_{q_{2}}]\) where \((\lambda_{p_{1}},\mu_{q_{1}})\) and \((\lambda_{p_{2}},\mu_{q_{2}})\) are two disjoint components of S. Then there exists an increasing sequence {s m } with t m <s m <τ m such that L′(s m )<0 and \(L(s_{m})\in(\mu_{q_{2}},\lambda_{p_{1}})\cap S^{c}\) for all m. Hence, \(x_{i}'(s_{m})<0\) for all i and L′(s m )>0, a contradiction.

Thus, [λ j ,μ j ] and [λ k ,μ k ] belong to the same set [λ p ,μ q ], that is, \(\overline{L}\) and \(\underline{L}\) belong to [λ p ,μ q ], where (λ p ,μ q ) is a component of S.

If there does not exist γΓ, Γ={λ i , μ i :i=1,2,…,n}, s.t. \(\gamma\in(\underline{L},\overline{L})\), then there exists some r s.t. L(t)∈[λ r ,μ r ]⊂[λ p ,μ q ] for all large t. Then we have \(x_{r}'(t)>0\) for all large t. From the boundedness of x r , it follows that \(\lim_{t\to\infty}x_{r}(t)=x_{r}^{*}\geq0\) and lim t→∞ L(t)=λ r or μ r , a contradiction.

Thus, there exists some γΓ s.t. \(\gamma\in(\underline{L},\overline{L})\), let γ 1, γ 2 be two consecutive elements of Γ such that \(\gamma_{1}<\underline{L}<\gamma_{2}\). Since L(t) oscillates, there exists T 1<T 2 such that L(T 1)=L(T 2)=γ 2, γ 1<L(t)≤γ 2 for t∈[T 1,T 2] and L′(T 1)<0<L′(T 2). From

$$L'(t)=-L(t) \Biggl[\sum_{i=1}^{n}k_ix_i'(t) \Biggr], $$

we have

$$\sum_{i=1}^{n}k_ix_i'(T_1)>0> \sum_{i=1}^{n}k_ix_i'(T_2). $$

We divide the above summation into two parts, one is \(x_{i}'(T_{j})<0\), i.e. g i (γ 2)<d i , the other is \(x_{i}'(T_{j})>0\), i.e. g i (γ 2)>d i . Therefore, we have

$$\sum_{g_i(\gamma_2)<d_i}k_ix_i'(T_1)+ \sum_{g_i(\gamma_2)>d_i}k_ix_i'(T_1) >\sum_{g_i(\gamma_2)<d_i}k_ix_i'(T_2)+ \sum_{g_i(\gamma_2)>d_i}k_ix_i'(T_2), $$

and

(15)

For the case g i (γ 2)<d i , i.e., γ 2∉(λ i ,μ i ), then L([T 1,T 2]) is disjoint from (λ i ,μ i ). Hence, \(x_{i}'(t)<0\) for t∈[T 1,T 2] and x i (T 2)<x i (T 1). For the case g i (γ 2)>d i , i.e., λ i <γ 2<μ i , we have L([T 1,T 2])⊂(λ i ,μ i ). Hence \(x_{i}'(t)>0\) for t∈[T 1,T 2] and x i (T 2)>x i (T 1). Thus,

a contradiction to (15). Hence, the theorem holds. □

Rights and permissions

Reprints and permissions

About this article

Cite this article

Hsu, SB., Lin, CJ., Hsieh, CH. et al. Dynamics of Phytoplankton Communities Under Photoinhibition. Bull Math Biol 75, 1207–1232 (2013). https://doi.org/10.1007/s11538-013-9852-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11538-013-9852-3

Keywords

Navigation