Citation: Ming Chen, Meng Fan, Xing Yuan, Huaiping Zhu. Effect of seasonal changing temperature on the growth of phytoplankton[J]. Mathematical Biosciences and Engineering, 2017, 14(5&6): 1091-1117. doi: 10.3934/mbe.2017057
[1] | Zhenyao Sun, Da Song, Meng Fan . Dynamics of a stoichiometric phytoplankton-zooplankton model with season-driven light intensity. Mathematical Biosciences and Engineering, 2024, 21(8): 6870-6897. doi: 10.3934/mbe.2024301 |
[2] | Lidan Liu, Meng Fan, Yun Kang . Effect of nutrient supply on cell size evolution of marine phytoplankton. Mathematical Biosciences and Engineering, 2023, 20(3): 4714-4740. doi: 10.3934/mbe.2023218 |
[3] | Mohammad A. Tabatabai, Wayne M. Eby, Sejong Bae, Karan P. Singh . A flexible multivariable model for Phytoplankton growth. Mathematical Biosciences and Engineering, 2013, 10(3): 913-923. doi: 10.3934/mbe.2013.10.913 |
[4] | Saswati Biswas, Pankaj Kumar Tiwari, Yun Kang, Samares Pal . Effects of zooplankton selectivity on phytoplankton in an ecosystem affected by free-viruses and environmental toxins. Mathematical Biosciences and Engineering, 2020, 17(2): 1272-1317. doi: 10.3934/mbe.2020065 |
[5] | Jean-Jacques Kengwoung-Keumo . Dynamics of two phytoplankton populations under predation. Mathematical Biosciences and Engineering, 2014, 11(6): 1319-1336. doi: 10.3934/mbe.2014.11.1319 |
[6] | He Liu, Chuanjun Dai, Hengguo Yu, Qing Guo, Jianbing Li, Aimin Hao, Jun Kikuchi, Min Zhao . Dynamics induced by environmental stochasticity in a phytoplankton-zooplankton system with toxic phytoplankton. Mathematical Biosciences and Engineering, 2021, 18(4): 4101-4126. doi: 10.3934/mbe.2021206 |
[7] | Juan Li, Yongzhong Song, Hui Wan, Huaiping Zhu . Dynamical analysis of a toxin-producing phytoplankton-zooplankton model with refuge. Mathematical Biosciences and Engineering, 2017, 14(2): 529-557. doi: 10.3934/mbe.2017032 |
[8] | Ann Nwankwo, Daniel Okuonghae . Mathematical assessment of the impact of different microclimate conditions on malaria transmission dynamics. Mathematical Biosciences and Engineering, 2019, 16(3): 1414-1444. doi: 10.3934/mbe.2019069 |
[9] | Ayako Suzuki, Hiroshi Nishiura . Seasonal transmission dynamics of varicella in Japan: The role of temperature and school holidays. Mathematical Biosciences and Engineering, 2023, 20(2): 4069-4081. doi: 10.3934/mbe.2023190 |
[10] | Hao Wang, Yang Kuang . Alternative models for cyclic lemming dynamics. Mathematical Biosciences and Engineering, 2007, 4(1): 85-99. doi: 10.3934/mbe.2007.4.85 |
Phytoplankton constitutes the bottom level of the aquatic food web and plays a fundamental role in the aquatic ecosystem. Phytoplankton bloom (PB) can assert severe negative consequences on ecosystems and social economics. In the past decades, there have been numerous field or experimental and also some mathematical modeling studies on the impact of some specific factors contributing to the growth of phytoplankton. Many researchers focused on the role of nutrients played on PB. Many studies show that eutrophication is the main reason leading to PB [9,18]. Johnson et al. [24] and Phillips et al. [30] stressed the importance of sediment phosphorus release in the restoration of very shallow lakes. Shukla et al. [32] modeled and analyzed PB in a lake caused by discharge of nutrients. Those studies can help to understand the contribution to PB from the different specific points. Nevertheless, most of the studies ignored the impact of temperature, which is thought to be a possible key triggering factor of PB.
There have been some studies emphasizing on the effect of constant temperature on the growth of phytoplankton [8,40]. Especially, Chen et al [5] presented a sound frame work on the effect of constant temperature on the growth of phytoplankton. They defined a helpful index called basic reproductive index to fully characterize the relationship between the dynamic evolution of phytoplankton and temperature.
The model considering the effect of constant temperature on the growth of phytoplankton is unsuitable for modeling PB that recurs seasonally. Then some aquatic ecologists have been shifting their attentions to seasonal variations of phytoplankton composition and has been attracted by concepts of periodicity, succession and response to environmental changes [33,38]. Some field or experimental studies (such as our field study in Fig. 2 and 3) suggested that the dominant variations in both phytoplankton composition and environmental variables are seasonal over a consistent annal cycle in lake and strongly related to the annual cycles of temperature [7,21,39]. Our goal of this paper is to develop modeling approaches for understanding the effect of seasonal changing temperature on the growth of phytoplankton and model the complex dynamics of seasonally recurring phytoplankton blooms.
The temperature plays important roles in many aspects during the process of phytoplankton metabolism. The impacts of temperature on the metabolism of phytoplankton are traditionally described by the
Located in the southern part of the Yangtze River delta, Lake Tai is the third largest freshwater lake in China with a surface area of about 2,338 square kilometers (about 902 square miles), an average depth of 2 meters, and maximum depth 3 meters. Lake Tai has been used as the important freshwater sources in the yangtze river delta on edge of Chinese eastern coast. Its water quality and environmental conditions have an important impact on the development of this area. In recent years, with the fast development of industry and agriculture process, waste water and domestic sewage without thorough treatment were discharged into Lake Tai Basin, then those pollutions gradually worsen the water quality and assert considerably negative impact on the ecological environment system of Lake Tai Basin. The water pollution of cyanobacteria blooms has become a severe problem in this area. At present, most of the studies were focused on Wuxi Meiliang bay of Lake Tai and the pollution in other typical areas around Lake Tai were less explored.
From March 2008 to April 2009, we have made an detailed field investigation of the water qualities and temperature variations around Lake Tai. The study was conducted in 12 selected monitor stations (i.e., MS) set in the typical areas around Lake Tai (Fig 1). We sampled the water, collected data, monitored and assessed the water quality once a month. The specimen water was sampled from
MS | Name | Longitude | Latitude |
| Jia Xing Canal | ||
| Wang Jiang Jing | ||
| Ping Wang Bridge | ||
| Hu Zhou Source | ||
| Xiao Mei Kou | ||
| New Port | ||
| Yi Xing Industry | ||
| Zhou Tie Sewage | ||
| Yi Xing Rv/Lake | ||
| Wu Xi Mei Liang | ||
| Wu Xi Lake | ||
| Su Zhou Lake Tai |
MS | | | | |
19.18 | -11.37 | 7.1527 | 0.364 | |
18.90 | -11.68 | 0.8611 | 0.242 |
The density of phytoplankton is expressed by the density of Chlorophyll-a. The time series and observed patterns of water temperature and Chlorophyll-a in the targeted monitor stations
The remainder of this paper is organized as follows. In section 2, we propose a time-varying nutrient-phytoplankton dynamic model including the time-varying effect of temperature on metabolism of phytoplankton. In section 3, an ecological reproductive index is proposed and the existence and stability of periodic solutions are proved by coincidence degree theory and Lyapunov direct method. Section 4 carries out numerical simulations to expound both the long term cyclic dynamics and the short term effect of daily temperature on the growth of phytoplankton. The filed application is then carried out for Lake Tai to validate our model. The paper ends with conclusion and some interesting discussion.
In this section, following the clues and framework in [5], we formulate a dynamical model governed by non-autonomous ordinary differential equations for a lake with inflow and outflow of nutrient and study the impact of time-varying temperature on the dynamic evolution of both the phytoplankton and the nutrient. We assume that the lake is a physically homogeneous shallow lake or the well-mixed surface layer of a stratified water column, which thoroughly mixed at all times, so that there are no spatial gradient of concentrations.
The forced phytoplankton-nutrient compartment model consists of two state variables. Let
The loss of phytoplankton due to natural mortality is assumed to be proportional to its density
{dAdt=μoptf(P)A⏟growth−mA⏟natural mortality−bA2⏟crowding loss,dPdt=aPs⏟input−Upf(P)A⏟uptake−aP⏟dilution loss. | (1) |
Next, we discuss the mathematical formulation of the components of (1), in particular, we will expound their dependence on the temperature.
Nutrient-limited functional response. The nutrient-limited functional response
f(P)∈C2, f(0)=0, f′(P)>0, lim | (2) |
The most typical and common used is the so-called Hill function
f_n(P)=\frac{P^n}{H_P^n+P^n}, | (3) |
which incorporates the Holling family of functional responses [22] as special cases, where
f_3(P)=\frac{P^{r_1}}{P^{r_1}+H_2P^{r_2}+H_1}, |
where
Temperature is a key environmental factor for the lake system and asserts essential impacts on the nutrient uptake, growth and death of phytoplankton, the nutrient releasing from particulate nutrients, and so on. Next, we shall deliberately discuss these metabolism progresses respectively.
Temperature-dependent uptake and growth. The nutrient uptake and growth of phytoplankton are heavily affected by the variation of temperature. There are several model candidates depicting the effect of temperature such as linear model [1,34], Michaelis-Menten-Monod saturating model [34], optimal exponential model [3,40] and Arrhenius model [17,19]. In this study, we adopt the so-called cardinal temperature model with inflexion (CTMI) proposed by Rosso et al [31], which is validated for a large range of species. As we all known, phytoplankton grows much better under more suitable temperature condition and halts growth under extremely low or high temperature. Bernard [2] shows that the CTMI can accurately represent the growth of microalgae. The CTMI includes four key parameters and three of them are cardinal temperatures with a biological significance, which makes the model rather straightforward to calibrate and facilitates the estimation with experimental data. It predicts the growth rate for temperatures between
g(T)= \begin{cases} 0,~~T<T_{min},\\ \phi(T),~~T_{min}<T<T_{max}, \\ 0,~~T>T_{max}, \end{cases} | (4) |
with
\phi(T)=\frac{(T-T_{max})(T-T_{min})^2}{(T_{opt}-T_{min})[(T_{opt}-T_{min})(T-T_{opt}) - (T_{opt}-T_{max})(T_{opt}+T_{min}-2T)]}. |
Here,
Temperature-dependent mortality. Water temperature essentially affects the natural mortality of phytoplankton. Higher temperature may lead to higher death rate [25]. Ding [8] adopted an exponential model to address the death rate with respect to water temperature, which reads
m(T)=m_0+K_1\theta_1^{T-T_{ref}}, |
where
Temperature-dependent nutrient release. We assume that there is an external source of nutrients (say phosphorus) flowing into the lake and the densities write as
Time-varying temperature. In real world application, the environmental factors and resources vary or fluctuate in time. The environmental fluctuations are likely to have a significant effect on the dynamic evolution of aquatic ecosystems. In order to study the effect of time-varying temperature forcing phytoplankton,
Consider the specific case that the temperature is assumed to be varying periodically, i.e.,
T(t)=T_{mean}+\alpha \sin(\Omega t+\varphi), | (5) |
where
To sum up, the objective phytoplankton-nutrient system takes the following form
\left\{ \begin{split} &\frac{dA}{dt}=\mu_{opt} g(T(t))f(P)A-m(T(t))A-bA^2,\\ &\frac{dP}{dt}=aP_s(T(t))-U_pg(T(t))f(P)A-aP. \end{split}\right. | (6) |
The description, unit, default value (when available), and reference resource for each parameter are summarized in Table 2.
Par. | Description | Value | Unit | Ref. |
maximum nutrient uptake coefficient | | | [40] | |
| maximum growth rate of phytoplankton | | | [2] |
| optimal water temperature | | | [2] |
| minimum water temperature | | | [2] |
| maximum water temperature | | | [2] |
| reference water temperature | | | [8] |
| natural mortality rate of phytoplankton | | | [8] |
| Respiration rate of phytoplankton | | | [8] |
| temperature dependence coefficient | | | [8] |
| temperature dependence coefficient | | | [26] |
| temperature dependence coefficient | | | [26] |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| density of total input soluble phosphorus | | | Defaulted |
| density of total input solid phosphorus | | | Defaulted |
| density of solid phosphorus in bottom sediments | | | Defaulted |
| dilution rate | | | Defaulted |
| density limiting coefficient | | | Defaulted |
In this section, we explore the dynamics of (6) and present some results on the positive invariance, the existence and globally asymptotic stability of boundary and internal cyclic dynamics. In order to facilitate the following discussion, write
g(t):=g(T(t)),~~m(t):=m(T(t)),~~P_s(t):=P_s(T(t)), |
which are assumed to be continuous and nonnegative.
For a bounded continuous function
h^u:=\mathop {\sup }\limits_{t\in R}h(t),~~h^l:=\mathop {\inf }\limits_{t\in R}h(t). |
Theorem 3.1. If
\Gamma:=\{(A,P)\in R^2\mid~m_1\leq A\leq M_1,~m_2\leq P\leq M_2 \} | (7) |
is positively invariant with respect to (6), where
\begin{align*} M_1:=\frac{\mu_{opt}g^u-m^l}{b},~~~~~~~~ M_2:=P_s^u, \end{align*} |
\begin{split} m_1:=\frac{\mu_{opt}g^lf(m_2)-m^u}{b},~ m_2:=\frac{aP_s^l-U_pg^uM_1}{a}.\\ \end{split} | (8) |
Proof. Let
A'(t)\leq bA(t)[\frac{\mu_{opt}g^u-m^l}{b}-A(t)]=bA(t)[M_1-A(t)], ~t\geq t_0. | (9) |
A standard comparison argument shows that
0<A(t_0)\leq M_1~\Rightarrow~ A(t)\leq M_1, t\geq t_0, |
which, together with the nutrient equation in (6), produces
P'(t)\leq a(P_s^u-P(t))=a(M_2-P(t)), ~t\geq t_0, | (10) |
and hence
0<P(t_0)\leq M_2~\Rightarrow~ P(t)\leq M_2, ~t\geq t_0. |
Similarly, the nutrient equation in (6) yields
P'(t)\geq a[\frac{aP_s^l-U_pg^uM_1}{a}-P(t)]=a[m_2-P(t)],~ t\geq t_0, | (11) |
whence
P(t_0)\geq m_2 ~\Rightarrow~P(t)\geq m_2,~t\geq t_0. |
Moreover, by the phytoplankton equation of system (6), we have
A'(t)\geq bA(t)[\frac{\mu_{opt}g^lf(m_2)-m^u}{b}-A(t)]=bA(t)[m_1-A(t)],~t\geq t_0, | (12) |
which implies
A(t_0)\geq m_1~\Rightarrow~A(t)\geq m_1,~t\geq t_0. |
Therefore,
Corollary 1. Let
Theorem 3.2. If
In the rest of Section 3,
g(t)=g(t+\omega),~~m(t)=m(t+\omega),~~P_s(t)=P_s(t+\omega). |
In order to characterize the population dynamics of phytoplankton, motivated by the computational formula of a threshold parameter for periodic compartmental epidemic models given by [37], here we introduce the ecological reproductive index
P_*(t)=(e^{a\omega}-1)^{-1}\int_t^{t+\omega}aP_s(s)e^{-a(t-s)}ds | (13) |
is the unique positive periodic solution of the following system
\frac{dP}{dt}=aP_s(t)-aP. | (14) |
Let
\frac{dE}{dt}=\mathcal{F}(t,E)-\mathcal{V}(t,E), | (15) |
where
\mathcal{F}(t,E)=\begin{pmatrix}\mu_{opt} g(t)f(P))A\\0\end{pmatrix},~ \mathcal{V^-}=\begin{pmatrix}m(t)A+bA^2\\aP+U_Pg(t)f(P)A\end{pmatrix},~ \mathcal{V^+}=\begin{pmatrix}0\\aP_s(t)\end{pmatrix}. | (16) |
By carrying out arguments similar to those in [10,Lemma 1], one has
D_E\mathcal{F}(t,E_0)=\begin{pmatrix}F(t) & 0 \\0 & 0\end{pmatrix},~ D_E\mathcal{V}(t,E_0)=\begin{pmatrix}V(t) & 0 \\J(t) & a\end{pmatrix}, | (17) |
where
\frac{dA}{dt}=\mu_{opt} g(t)f(P_*(t))A-m(t)A=(F(t)-V(t))A. | (18) |
Denote by
\frac{dy(t)}{dt}=-V(t)y, | (19) |
that is, for each
\frac{d}{dt}Y(t,s)=-V(t)Y(t,s),~\forall t\geq s,~ Y(s,s)=1. | (20) |
Hence, the monodromy matrix
Assume that
\psi(t):=\int_{-\infty}^t Y(t,s)F(s)\phi(s)ds=\int_0^\infty Y(t,t-a)F(t-a)\phi(t-a)da |
is the distribution of the accumulative new individuals at time
Let
(L\phi)(t)=\int_0^\infty Y(t,t-a)F(t-a)\phi(t-a)da,~\forall t\in R, ~\phi\in C_{\omega}, | (21) |
that is
(L\phi)(t)=\int_0^\infty\mu_{opt} g(t-a)f(P_*(t-a))e^{-\int_{t-a}^t m(s)ds}\phi(t-a)da, ~\forall t\in R, ~\phi\in C_{\omega}. |
From [36] and [37], it follows that the ecological reproductive index
In order to formulate the explicit expression of
\frac{dw}{dt}=[-V(t)+\frac{F(t)}{\sigma}]w,~~t\in R_+ | (22) |
with
Lemma 3.3. ([37]
(ⅰ) If
(ⅱ) If
(ⅲ)
Lemma 3.4. ([37]
(ⅰ)
(ⅱ)
(ⅲ)
Let
W(t,s,\sigma)=\Phi_{(F/\sigma)-V}(t),~~\Phi_{F-V}(t)=W(t,0,1),~~t\geq 0. |
By Theorem 2.1 and 2.2 in [37], the ecological reproductive index can be also defined as
R_0=\left(\displaystyle\int_0^\omega \mu_{opt} g(t)f(P_*(t))dt\right)\left(\displaystyle\int_0^\omega m(t)dt\right)^{-1}. |
The ecological reproductive index
In this subsection, we explore the boundary cyclic dynamics of (6) by establishing the existence and globally asymptotic stability of the boundary
Definition 3.5. A bounded non-negative solution
\mathop {\lim }\limits_{t\rightarrow+\infty}(|x(t)-\hat{x}(t)|+|y(t)-\hat{y}(t)|)=0. |
Lemma 3.6. ([41]
Theorem 3.7. (6) always has a boundary
Proof. By (6), one has
\frac{d\bar{P}}{dt}=aP_s(t)-a\bar{P}. | (23) |
It is trivial to show that
\frac{dA}{dt}\leq[\mu_{opt} g(t)f(P_*(t)+\epsilon)-m(t)]A:=(F(t,\epsilon)-V(t))A, | (24) |
where
In the following, we investigate the internal cyclic dynamics of (6) by establishing some sufficient criteria for the existence and GAS of positive periodic solution of (6). For the reader's convenience, we shall summarize below a few concepts and results from [16] that will be essential for the following discussion.
Let
Lemma 3.8 (Continuation theorem[16]). Let
(a) for each
(b)
Theorem 3.9. Assume that
\hat{R_0}=\frac{\mu_{opt}\bar{g}f(K_2)}{\bar{m}}>1, | (25) |
then (6) has at least one strictly positive
F(x)=ax+U_p\bar{g}f(x)e^{K_1},~~K_1=\ln\frac{\mu_{opt}\bar{g}}{b}+2\omega \mu_{opt}\bar{g},~~K_2=F^{-1}(a\bar{P_s})-2\omega a\bar{P_s}>0. |
Proof. Considering the biological significance of (6), we specify
\overline{g}=\frac{1}{\omega}\int_0^{\omega}g(t)dt,~~\bar{m}=\frac{1}{\omega}\int_0^{\omega}m(t)dt,~~\bar{P_s}=\frac{1}{\omega}\int_0^{\omega}P_s(t)dt. |
Let
\begin{split} &\frac{dx_1(t)}{dt}=\mu_{opt} g(t)f(x_2(t))-m(t)-b\exp(x_1(t)),\\ &\frac{dx_2(t)}{dt}=a(P_s(t) -x_2(t))-U_pg(t)f(x_2(t))\exp(x_1(t)). \end{split} | (26) |
Define
X=Z=\{x(t)=(x_1(t),x_2(t))^T\in C(R,R^2),x(t+\omega)=x(t)\}, |
and
Nx= \left(\begin{array}{ccc} \mu_{opt} g(t)f(x_2(t))-m(t)-b\exp(x_1(t)) \\ a(P_s(t) -x_2(t))-U_pg(t)f(x_2(t))\exp(x_1(t)) \end{array}\right),~~x\in X, |
Lx=\frac{dx(t)}{dt},~Px=\frac{1}{\omega}\int_0^{\omega}x(t)dt,~x\in X;~~ Qz=\frac{1}{\omega}\int_0^{\omega}z(t)dt, z\in Z. |
Then, it is not difficult to show that
Now we are at the right position to search for an appropriate open bounded subset
\begin{split} &\frac{dx_1(t)}{dt}=\lambda\{\mu_{opt} g(t)f(x_2(t))-m(t)-b\exp(x_1(t))\},\\ &\frac{dx_2(t)}{dt}=\lambda\{a(P_s(t) -x_2(t))-U_pg(t)f(x_2(t))\exp(x_1(t))\}. \end{split} | (27) |
Let
\int_0^\omega{\{\mu_{opt} g(t)f(x_2(t))-m(t)-b\exp(x_1(t))}\}dt=0, | (28) |
\int_0^\omega\{a(P_s(t) -x_2(t))-U_pg(t)f(x_2(t))\exp(x_1(t))\}dt=0, | (29) |
then
\int_0^\omega\{m(t)+b\exp(x_1(t))\}dt=\int_0^\omega \mu_{opt} g(t)f(x_2(t))dt\leq \int_0^\omega \mu_{opt} g(t)dt=\omega \mu_{opt}\bar{g}, | (30) |
\int_0^\omega{aP_s(t)dt=\int_0^\omega \{ax_2(t)+U_pg(t)f(x_2(t))\exp(x_1(t))}\}dt. | (31) |
From (27) - (31), it follows that
\begin{split} \displaystyle\int_0^\omega|\dot{x}_1(t)|dt & = \lambda\displaystyle\int_0^\omega{|\mu_{opt} g(t)f(x_2(t))-m(t)-b \exp(x_1(t))}|dt\\ & < \displaystyle\int_0^\omega|\mu_{opt} g(t)f(x_2(t))|dt+\displaystyle\int_0^\omega |m(t)+b \exp(x_1(t))|dt=2\omega\mu_{opt}\overline{g}, \end{split} |
and
\begin{split} \displaystyle\int_0^\omega|\dot{x}_2(t)|dt &= \lambda\displaystyle\int_0^\omega{|a(P_s(t) -x_2(t))-U_pg(t)f(x_2(t))\exp(x_1(t))}|dt\\ &< \displaystyle\int_0^\omega|aP_s(t)|dt+\displaystyle\int_0^\omega |ax_2(t)+U_pg(t)f(x_2(t))\exp(x_1(t))|dt<2\omega a\overline{P_s}. \end{split} |
Since
x_i(\xi_i)=\mathop {\min }\limits_{t\in[0,\omega]}x_i(t),~~ x_i(\eta_i)=\mathop {\max }\limits_{t\in[0,\omega]}x_i(t), i=1,2. |
By (28), one has
b\omega \exp(x_1(\xi_1))\leq\int_0^\omega\{m(t)+b\exp(x_1(t))\}dt\leq \int_0^\omega \mu_{opt} g(t)dt=\omega \mu_{opt}\bar{g}, |
that is
x_1(t)\leq x_1(\xi_1)+\int_0^\omega|\dot{x}_1(t)|dt\leq\ln\frac{\mu_{opt}\bar{g}}{b}+2\omega \mu_{opt}\bar{g}:=K_1. |
By (29), one has
then
x_2(t)\leq x_2(\xi_2)+\int_0^\omega|\dot{x}_2(t)|dt\leq \bar{P_s}+2\omega a\bar{P_s}. |
Moreover, (31) implies
\int_0^\omega{aP_s(t)dt\leq \int_0^\omega \{ax_2(\eta_2)+U_pg(t)f(x_2(\eta_2))\exp(K_1)}\}dt, |
then
a\bar{P_s}\leq ax_2(\eta_2)+U_p\bar{g}f(x_2(\eta_2))\exp(K_1):=F(x_2(\eta_2)). |
Since
x_2(\eta_2)\geq F^{-1}(a\bar{P_s})>0, |
whence
x_2(t)\geq x_2(\eta_2)-\int_0^\omega|\dot{x}_2(t)|dt\geq F^{-1}(a\bar{P_s})-2\omega a\bar{P_s}=K_2>0. |
Therefore,
\int_0^\omega \mu_{opt} g(t)f(K_2)dt\leq (\bar{m}+b\exp(x_1(\eta_1)))\omega, |
then
x_1(\eta_1)>\ln\frac{\mu_{opt}\bar{g}f(K_2)-\bar{m}}{b}, |
thus
x_1(t)\geq x_1(\eta_1)-\int_0^\omega|\dot{x}_1(t)|dt\geq\ln\frac{\mu_{opt}\bar{g}f(K_2)-\bar{m}}{b}-2\omega \mu_{opt}\bar{g}. |
Therefore,
\mathop {\max }\limits_{t\in[0,\omega]}|x_1(t)|\leq\max\{|\ln \frac{\mu_{opt}\bar{g}}{b}+2\omega \mu_{opt}\bar{g}|, |\ln\frac{\mu_{opt}\bar{g}f(K_2)-\bar{m}}{b}-2\omega \mu_{opt}\bar{g}|\}:=K_4. |
Clearly,
\begin{align*} \mu_{opt} \bar{g}f(v_2)-\bar{m}-bv_1&=0,~~~ a(\bar{P_s}-v_2)-U_p\bar{g}f(v_2)v_1=0. \end{align*} |
has a unique solution
QNx= \left(\begin{array}{ccc} \mu_{opt} \bar{g}f(x_2)-\bar{m}-b\exp(x_1) \\ a(\bar{P_s} -x_2)-U_p\bar{g}f(x_2)\exp(x_1) \end{array}\right)\neq0. |
Furthermore, in view of the assumptions in Theorem 3.9, direct calculations show that
Next, we go ahead with investigating the globally asymptotic stability of the positive
Lemma 3.10. Let
Theorem 3.11. Let
R_1:=\frac{\mu_{opt}g^lf(m_2)}{m^u}>1,~~aP_s^l>U_pg^uM_1, |
\mathop {\inf }\limits_{t\in R}\{b-U_pg(t)f(M_2)\}>0,~~\inf\limits_{t\in R}\{U_pg(t)m_1-\mu_{opt}g(t)\}>0, | (32) |
where
Proof. Let
Consider the Lyapunov function defined by
V(t)=|\ln\{A(t)\}-\ln\{A^*(t)\}|+|P(t)-P^*(t)|. |
A direct calculation of the right derivative of
\begin{split} V'(t)=& -a|P(t)-P^*(t)|-b|A(t)-A^*(t)|\\ &+sgn\{P(t)-P^*(t)\}U_pg(t)(-f(P(t))A(t)+f(P^*(t))A^*(t))\\ &+sgn\{A(t)-A^*(t)\}\mu_{opt}g(t)(f(P(t))-f(P^*(t)))\\ \leq & -a|P(t)-P^*(t)|-b|A(t)-A^*(t)|+U_pg(t)f(P(t))|A(t)-A^*(t)|\\ &-U_pg(t)A^*(t)|f(P(t))-f(P^*(t))|+\mu_{opt}g(t)|f(P(t))-f(P^*(t))|\\ \leq & -a|P(t)-P^*(t)|-(b-U_pg(t)f(M_2))|A(t)-A^*(t)|\\ &-(U_pg(t)m_1-\mu_{opt}g(t))|f(P(t))-f(P^*(t))|. \end{split} |
Note that there are some terms containing
f(P^*(t))A^*(t)-f(P(t))A(t)=A^*(t)(f(P^*(t))-f(P(t))+f(P(t))(A^*(t)-A(t)), |
from (32), it follows that there exists a positive constant
D^+V(t)\leq -\phi[|A(t)-A^*(t)|+|P(t)-P^*(t)|+|f(P(t))-f(P^*(t))|],~t\geq t_0+T_1. | (33) |
Integrating both sides of (33) from
\begin{split} V(t)+\phi\int_{t_0+T_1}^t[|A(s)-A^*(s)|&+|P(s)-P^*(s)|+|f(P(s))-f(P^*(s))|]ds\\ &\leq V(t_0+T_1)<+\infty. \end{split} |
Then
\begin{split} \int_{t_0+T_1}^t[|A(s)&-A^*(s)|+|P(s)-P^*(s)|+|f(P(s))-f(P^*(s))|]ds\\ &\leq V(t_0+T_1)/\phi<+\infty,~t\geq t_0+T_1, \end{split} |
Therefore,
The boundedness of
\mathop {\lim }\limits_{t\rightarrow+\infty}(|A(t)-A^*(t)|+|P(t)-P^*(t)|)=0. |
The proof is complete.
In summary, we have established some sufficient criteria for the boundary and internal cyclic dynamics of (6). Fig. 5(a) and (b) illustrate such scenarios, where the parameter values are deliberately specified such that the criteria in Theorem 3.7 and Theorem 3.11 are satisfied, respectively. In addition, it is not difficult to show that
R_0\geq\frac{\displaystyle\int_0^\omega \mu_{opt} g^lf(m_2)dt}{\displaystyle\int_0^\omega m^udt}=R_1, ~~ i.e., ~~ R_1>1 ~\mathrm{implies} ~R_0>1. |
The criteria established above are of sufficient type and have some room for further improvement. Fig. 5(c) illustrates such an example, where the conditions in Theorems 3.11 are not satisfied (i.e.,
In this section, based on the above theoretical analysis, by using the numerical simulations and field applications, we further investigate the ecological impact and consequence of temperature variation and the system's corresponding response.
The general form of (6) and the definition of
Fig. 8 elaborates the relationship between the amplitude of phytoplankton cyclic oscillation and the cyclic forcing of water temperature. The figure shows the amplitude of the cyclic oscillation of phytoplankton increases first and then decreases with the increasing of the strength of cyclic forcing of water temperature. The relationship between the amplitude of temperature and the phytoplankton amplitude is not simply monotonously consistent. The amplitude of phytoplankton oscillation achieves a maximum value at some intensity of forcing temperature.
Based on the analysis of 47 years observation data of
T(t)=T_{mean}+\alpha\sin(\Omega t+\varphi). | (34) |
Since the forcing is assumed to be monthly and
According to studies on zooplankton composition and feeding rates of crustacean in Lake Tai [6,7], the zooplankton there could never reach values high enough to control Microcystis production. Moreover, the dominant fish population in Lake Tai was zooplankton-feeding fish and phytophagous fish abundance was insignificant to control the Microcystis bloom. That is to say, phytoplankton in Lake Tai should therefore not be controlled by the grazing of either zooplankton or fish. Therefore, the top-down control can be reasonably ignored and the bottom-up formulation is adequate as a general description in mathematical modeling. Therefore, the nutrient-phytoplankton model established in this study can be applied to model and to predict phytoplankton dynamics in Lake Tai.
In Lake Tai, temperature plays an important role in the phytoplankton composition and the dominant variations in phytoplankton composition are seasonal and are strongly related to the annual cycles of temperature [6,7]. It is reasonable to assume that the annual cycles of growth, reproduction, and senescence of phytoplankton are finely tuned to the annual climate cycle having a period of one year. Cyclic behavior of phytoplankton has been shown to occur in both the field and the laboratory cultures and reconciling the predictions of mathematical models with the nonequilibrium dynamics of experimental communities has proved to be challenging [15].
As example, we only fit our model to monitor set
We treat Lake Tai as a whole and consider the average of the collected data in each monitoring station. Fig. 11 shows the average temperature and chlorophyll-a. The numerical simulation shows that the prediction curve well matches the changes of chlorophyll-a biomass. It is not difficult to observe that the fitness in Fig. 10 is better than that in Fig. 11, which is due to the fact that the whole lake contains more uncertain impact factors. It can be more accurate to predict a portion than a whole. Despite this, our model can simulate the trend of changes of phytoplankton biomass in the whole lake.
In order to better assess the model performance, adaptability, and to analyze the impact and consequences of temperature on model dynamics, below we present numerical simulations to show short time responses to the changing temperature.
We consider the water temperature being constant with an outbreak. In general, the temperature can be simulated by the following convenient and common formulation
T(t)=\max(T_{mean},T_{high}-\delta(t-5)^2),~~t\in[0,15], |
where
In all the four cases in Fig. 12, the density of phytoplankton also shows an outbreak under the corresponding temperature forcing. One can study Fig. 12 by comparing (a) with (b), (c), and (d). The wider 'open mouth' of temperature variation leads to the wider 'open mouth' of the phytoplankton outbreak (Fig. 12(a) and (b)). In addition to this finding, we also find that the peak value of phytoplankton in (b) is higher than that in (a). That is to say, longer period high temperature causes the phytoplankton stay in higher value for long time. The magnitude of the temperature variation is also important. The bigger the temperature changes, the stronger the phytoplankton fluctuates (Fig. 12(a) and (c)). Even though with the same magnitude of the temperature variation, the mean temperature also plays an important role. One can consider the case (a) as summer and case (d) as winter. The temperature with the same magnitude causes the different results (Fig. 12(a) and (d)). That is to say, phytoplankton bloom is more likely to happen during the summer. The four cases in Fig. 12 also illustrate that the peak of the density of phytoplankton occurs later than the time maximum temperature occurs (Fig. 12(a)). That is to say, temperature may have a lag effect on growth of phytoplankton. The time of maximum of phytoplankton falls in the range where
We would like to point out that the phytoplankton with different optimal temperatures can have different responses to even the same temperature forcing. The time series of phytoplankton with optimal temperature
Because of ignoring the time-varying nature of growth rate and environmental biomass of phytoplankton, the traditional\break autonomous model could not well explain why phytoplankton usually occur in high temperature seasons [28]. In this study, in order to investigate the effect of seasonal temperature on the growth of phytoplankton, by incorporating the time-varying feature of temperature, we propose a new time-dependent bottom-up nutrient-phytoplankton interacting model. Our model provides a reasonable explanation to understand the mechanism of phytoplankton bloom.
Our model has made some improvements of the traditional models. Effect of temperature on metabolism of phytoplankton like growth, nutrient uptake, natural mortality, and nutrient releasing from sediments are analyzed systematically and comprehensively. Especially, a more realistic model named CTMI is incorporated into our model. CTMI is determined by three parameters (
Certainly, in order to obtain a desired model to well expound the mechanism of temperature working on the growth of phytoplankton or even the triggering mechanism of PB, we still has a very long way to go. Freund [14] found that an average blooms are correlated with rapid upward temperature fluctuations and speculate on their possible role as trigger mechanisms. Daily average temperature may have more immediate effect on the phytoplankton growth and bloom. The effect of daily temperature on phytoplankton bloom should be studied intensively. To some extend, our model is still an idealized one, for example, here we view temperature as a periodic function of time, yet in the real world, the water temperature changes in some random way [28]. Hence, it is more realistic to take into account the stochastic noise in the equation predicting water temperature in our model and to implement further theoretical or numerical analysis.
Our study provides a preliminary explanation of how temperature contributes to the recurrent outbreaks of PB. To further understand the mechanism of PB, one should consider other abiotic and biotic factors that are responsible for PB. The light intensity, precipitation, ecological stoichiometry of growth-limiting nutrient (e.g., the ratio of nitrogen to phosphorus, expressed as N:P) etc., may also have significant effect on the metabolism of phytoplankton. Although our climate-driven bottom-up nutrient-phytoplankton model sheds some new light on the mechanism of phytoplankton dynamics, the alternative regulation of phytoplankton abundance can come from other biotic processes such as the important top-down control. The food web interactions also have a profound effect on the population dynamics of phytoplankton bloom species and it is left for our future effort.
[1] | [ G. Ahlgren, Temperature functions in biology and their application to algal growth constants, Oikos, 49 (1987): 177-190. |
[2] | [ O. Bernard,B. Remond, Validation of a simple model accounting for light and temperature effect on microalgal growth, Bioresource Technology, 123 (2012): 520-527. |
[3] | [ R. Bouterfas,M. Belkoura,A. Dauta, Light and temperature effects on the growth rate of three freshwater [2pt] algae isolated from a eutrophic lake, Hydrobiologia, 489 (2002): 207-217. |
[4] | [ C. Butterwick,S.I. Heaney,J. F. Talling, Diversity in the influence of temperature on the growth rates of freshwater algae, and its ecological relevance, Freshwater Biol, 50 (2005): 291-300. |
[5] | [ M. Chen,M. Fan,R. Liu,X. Yuan,H. P. Zhu, The dynamics of temperature and light on the growth of phytoplankton, J. Theor. Biol., 385 (2015): 8-19. |
[6] | [ W. Chen,A. Nauwerck, A note on composition and feeding of the crustacean zooplankton of Lake Taihu, Jiangsu Province, China, Limnologica, 26 (1996): 275-280. |
[7] | [ Y. W. Chen,B. Q. Qin,K. Teubner,M. T. Dokulil, Long-term dynamics of phytoplankton assemblages: Microcystis-domination in Lake Taihu, a large shallow lake in China, J. Plankton Res., 25 (2003): 445-453. |
[8] | [ L. Ding,Y. Pang,L. Li, Simulation study on algal dynamics under different hydrodynamic conditions, Acta Ecologica Sinica, 25 (2005): 1863-1868. |
[9] | [ X. H. Dong,H. Bennion,R. Battarbee,X. D. Yang,E. F. Liu, Tracking eutrophication in Taihu Lake using the diatom record: potential and problems, J. Paleolimnol, 40 (2008): 413-429. |
[10] | [ P. Driessche,W. James, Reproduction numbers and sub-threshold endemic equilibria for compartmental models of disease transmission, Math. Biosci., 180 (2002): 29-48. |
[11] | [ J. J. Elser,K. Acharya,M. Kyle,J. Cotner,W. Makino,T. Markow,T. Watts,S. Hobbie,W. Fagan,J. Schade,J. Hood,R. W. Sterner, Growth rate-stoichiometry couplings in diverse biota, Ecology Letters, 6 (2003): 936-943. |
[12] | [ M. Fan,Q. Wang,X. Zou, Dynamics of a non-autonomous ratio-dependent predator-prey system, Proceedings of the Royal Society of Edinburgh: Section A Mathematics, 133 (2003): 97-118. |
[13] | [ P. J. S. Franks, NPZ models of plankton dynamics: Their construction, coupling to physics, and application, J. Oceanog, 58 (2002): 379-387. |
[14] | [ J. A. Freund,S. Mieruch,B. Scholze,K. Wiltshire,U. Feudel, Bloom dynamics in a seasonally forced phytoplankton zooooplankton model: trigger mechanisms and timing effects, Ecol. Complex, 3 (2006): 129-139. |
[15] | [ G. F. Fussmann,S. P. Ellner,K. W. Shertzer,N. G. Hairston, Crossing the hopf bifurcation in a live predator-prey system, Science, 290 (2000): 1358-1360. |
[16] | [ R. E. Gaines and J. L. Mawhin, Coincidence Degree and Nonlinear Differential Equations, Springer-Verlag, Berlin, 1977. |
[17] | [ R. J. Geider,H. L. Maclntyre,T. M. Kana, A dynamic regulatory model of phytoplanktonic acclimation to light, nutrients, and temperature, Limnol. Oceanogr, 43 (1998): 679-694. |
[18] | [ P. M. Glibert, Eutrophication and Harmful Algal Blooms: A Complex Global Issue, Examples from the Arabian Seas including Kuwait Bay, and an Introduction to the Global Ecology and Oceanography of Harmful Algal Blooms (GEOHAB) Programme, Int. J. Oceans and Oceanography, 2 (2007): 157-169. |
[19] | [ J. C. Goldman,J. C. Edward, A kinetic approach to the effect of temperature on algal growth, Limnol. Oceanogr., 19 (1974): 756-766. |
[20] | [ G. M. Grimaud,V. L. Guennec,S. D. Ayata,F. Mariret,A. Schiandra,O. Bernard, Modelling the effect of temperature on phytoplankton growth across the global ocean, IFAC-PapersOnLine, 48 (2015): 228-233. |
[21] | [ J. P. Grover,T. H. Chrzanowski, Seasonal dynamics of phytoplankton in two warm temperate reservoirs: association of taxonomic composition with temperature, J. Plankt. Res, 28 (2006): 1-17. |
[22] | [ C. S. Holling, The functional response of predators to prey density and its role in mimicry and population regulation, Memoirs of the Entomological Society of Canada, 97 (1965): 5-60. |
[23] | [ A. Huppert,B. Blasius,R. Olinky,L. Stone, A model for seasonal phytoplankton blooms, J. Theor. Biol., 236 (2005): 276-290. |
[24] | [ K. S. Johnson,F. P. Chavez,G. E. Friederich, Continental-shelf sediment as a primary source of iron for coastal phytoplankton, Nature, 398 (1999): 697-700. |
[25] | [ R. I. Jones, The importance of temperature conditioning to the respiration of natural phytoplankton communities, British Phycological Journal, 12 (2007): 277-285. |
[26] | [ S. E. Jorgensen and G. Bendoricchio, Fundamentals of Ecological Modelling, Elsevier, 2001. |
[27] | [ I. Loladze,Y. Kuang,J. J. Elser, Stoichiometry in producer-grazer systems: Linking energy flow with element cycling, Bulletin of Mathematical Biology, 62 (2000): 1137-1162. |
[28] | [ J. H. Luo, Phytoplankton-zooplankton dynamics in periodic environments taking into account eutrophication, Math. Biosci., 245 (2013): 126-136. |
[29] | [ J. R. Moisan,T. A. Moisan,M. R. Abbott, Modelling the effect of temperature on the maximum growth rates of phytoplankton populations, Ecol. Model., 153 (2002): 197-215. |
[30] | [ G. Phillips,R. Jackson,C. Bennett,A. Chilvers, The importance of sediment phosphorus release in the restoration of very shallow lakes (The Norfolk Broads, England) and implications for biomanipulation, Hydrobiologia, 94 (1994): 445-456. |
[31] | [ L. Rosso,J. R. Lobry,J. P. Flandrois, An unexpected correlation between cardinal temperatures of microbial growth highlighted by a new model, J. Theor. Biol., 162 (1993): 447-463. |
[32] | [ J. B. Shukla,A. K. Misra,P. Chandra, Modeling and analysis of the algal bloom in a lake caused by discharge of nutrients, App. Math. Comput., 196 (2008): 782-790. |
[33] | [ S. J. Thackeray,I. D. Jones,S. C. Maberly, Long-term change in the phenology of spring phytoplankton: species-specific responses to nutrient enrichement and climatic changes, J. Ecol., 96 (2008): 523-535. |
[34] | [ D. Toro,M. Dominic,D. J. O'Connor,R. V. Thomann, A dynamic model of the phytoplankton population in the Sacramento San Joaquin Delta, Adv. Chem. Ser, 106 (1971): 131-180. |
[35] | [ C. L. Wang,W. Y. Pan,Y. Q. Han,X. Qian, Effect of global climate change on cyanobacteria bloom in taihu lake, China Environmental Science, 30 (2010): 822-828. |
[36] | [ F. B. Wang,S. B. Hsu,W. D. Wang, Dynamics of harmful algae with seasonal temperature variations in the cove-main lake, Discrete and Continuous Dynams. Systems -B, 21 (2016): 313-315. |
[37] | [ W. D. Wang,X. Q. Zhao, Threshold dynamics for compartmental epidemic models in periodic environments, J. Dyn. Differ. Equ., 20 (2008): 699-717. |
[38] | [ M. Winder,D. E. Schindler, Climate change uncouples trophic interactions in an aquatic ecosystem, Ecology, 85 (2004): 2100-2106. |
[39] | [ M. Winder,J. E. Cloern, The annual cycles of phytoplankton biomass, Philos. T. R. Soc. B, 365 (2010): 3215-3226. |
[40] | [ Q. J. Xu,B. Q. Qin,W. M. Chen,Y. W. Chen,G. Gao, Ecological simulation of algae growth in Taihu Lake, J. Lake Sci., 2 (2001): 149-157. |
[41] | [ F. Zhang,X. Q. Zhao, A periodic epidemic model in a patchy environment, J. Math. Anal. Appl., 325 (2007): 496-516. |
[42] | [ Q. Zhao,X. Lu, Parameter estimation in a three-dimensional marine ecosystem model using the adjoint technique, J. Marine Syst., 74 (2008): 443-452. |
[43] | [ X. Q. Zhao, Dynamical Systems in Population Biology, Springer-Verlag, New York, 2003. |
1. | S Karina, S Agustina, N Nurfadillah, A A Rafsanjani, J Heriantoni, Analysis of chlorophyll-a and phytoplankton abundance in Ujung Pancu Waters, Aceh Besar, Indonesia, 2021, 674, 1755-1307, 012079, 10.1088/1755-1315/674/1/012079 | |
2. | Da Song, Meng Fan, Shihan Yan, Meng Liu, Dynamics of a nutrient-phytoplankton model with random phytoplankton mortality, 2020, 488, 00225193, 110119, 10.1016/j.jtbi.2019.110119 | |
3. | Ming Chen, Meng Fan, Xin Wang, EFFECT OF TEMPERATURE ON ADAPTIVE EVOLUTION OF PHYTOPLANKTON CELL SIZE, 2020, 10, 2156-907X, 2644, 10.11948/20200008 | |
4. | M D Sani, P A Wiradana, A Y Maharani, R E Mawli, A T Mukti, The dominance and proportions of plankton in Pacific white shrimp (Litopenaeus vannamei) ponds cultivated with the intensive system in Bulukumba Regency, South Sulawesi, Indonesia, 2022, 1036, 1755-1307, 012057, 10.1088/1755-1315/1036/1/012057 | |
5. | Nihal G. Shams El-Din, Mostafa M. El-Sheekh, Hala Y. El-Kassas, D. I. Essa, Basma A. El-Sherbiny, Biological indicators as tools for monitoring water quality of a hot spot area on the Egyptian Mediterranean Coast, 2022, 15, 1866-7511, 10.1007/s12517-022-10701-6 | |
6. | Michael Chapwanya, Phindile Dumani, Stationary and oscillatory patterns in microbial population under environmental stress, 2023, 03784754, 10.1016/j.matcom.2023.03.022 | |
7. | Michael Chapwanya, Phindile Dumani, Spatio-temporal dynamics of microbial population under nutrient-limiting conditions, 2023, 456, 00963003, 128138, 10.1016/j.amc.2023.128138 | |
8. | Tiancai Liao, Dynamical complexity driven by water temperature in a size-dependent phytoplankton-zooplankton model with environmental variability, 2023, 05779073, 10.1016/j.cjph.2023.11.025 | |
9. | Jing Xia, Yalin Bao, Yonghui Gao, Ji Li, The effects of temperature and sulfamethoxazole on the growth and photosynthetic characteristics of Phaeodactylum tricornutum, 2024, 200, 0025326X, 116122, 10.1016/j.marpolbul.2024.116122 | |
10. | Michael Chapwanya, Phindile Dumani, 2024, 793, 978-1-4704-7606-9, 75, 10.1090/conm/793/15879 | |
11. | Zaitao Liang, Fangfang Liao, Feng Wang, Lyapunov stability of the Basener–Ross system, 2024, 0026-9255, 10.1007/s00605-024-01990-y | |
12. | Zhenyao Sun, Da Song, Meng Fan, Dynamics of a stoichiometric phytoplankton-zooplankton model with season-driven light intensity, 2024, 21, 1551-0018, 6870, 10.3934/mbe.2024301 | |
13. | Zaitao Liang, Haining Zhu, Exploring the periodic behavior of a singular predator-prey system, 2024, 0003-889X, 10.1007/s00013-024-02074-x | |
14. | Shengmei Lyu, Man Hu, Yi Zhu, Zhimao Deng, Limin Duan, Ruizhong Gao, Guoqiang Wang, Response mechanisms of eukaryotic plankton community structure to complex environmental conditions in semi-arid river basins, China, 2025, 376, 03014797, 124527, 10.1016/j.jenvman.2025.124527 |
MS | Name | Longitude | Latitude |
| Jia Xing Canal | ||
| Wang Jiang Jing | ||
| Ping Wang Bridge | ||
| Hu Zhou Source | ||
| Xiao Mei Kou | ||
| New Port | ||
| Yi Xing Industry | ||
| Zhou Tie Sewage | ||
| Yi Xing Rv/Lake | ||
| Wu Xi Mei Liang | ||
| Wu Xi Lake | ||
| Su Zhou Lake Tai |
MS | | | | |
19.18 | -11.37 | 7.1527 | 0.364 | |
18.90 | -11.68 | 0.8611 | 0.242 |
Par. | Description | Value | Unit | Ref. |
maximum nutrient uptake coefficient | | | [40] | |
| maximum growth rate of phytoplankton | | | [2] |
| optimal water temperature | | | [2] |
| minimum water temperature | | | [2] |
| maximum water temperature | | | [2] |
| reference water temperature | | | [8] |
| natural mortality rate of phytoplankton | | | [8] |
| Respiration rate of phytoplankton | | | [8] |
| temperature dependence coefficient | | | [8] |
| temperature dependence coefficient | | | [26] |
| temperature dependence coefficient | | | [26] |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| density of total input soluble phosphorus | | | Defaulted |
| density of total input solid phosphorus | | | Defaulted |
| density of solid phosphorus in bottom sediments | | | Defaulted |
| dilution rate | | | Defaulted |
| density limiting coefficient | | | Defaulted |
MS | Name | Longitude | Latitude |
| Jia Xing Canal | ||
| Wang Jiang Jing | ||
| Ping Wang Bridge | ||
| Hu Zhou Source | ||
| Xiao Mei Kou | ||
| New Port | ||
| Yi Xing Industry | ||
| Zhou Tie Sewage | ||
| Yi Xing Rv/Lake | ||
| Wu Xi Mei Liang | ||
| Wu Xi Lake | ||
| Su Zhou Lake Tai |
MS | | | | |
19.18 | -11.37 | 7.1527 | 0.364 | |
18.90 | -11.68 | 0.8611 | 0.242 |
Par. | Description | Value | Unit | Ref. |
maximum nutrient uptake coefficient | | | [40] | |
| maximum growth rate of phytoplankton | | | [2] |
| optimal water temperature | | | [2] |
| minimum water temperature | | | [2] |
| maximum water temperature | | | [2] |
| reference water temperature | | | [8] |
| natural mortality rate of phytoplankton | | | [8] |
| Respiration rate of phytoplankton | | | [8] |
| temperature dependence coefficient | | | [8] |
| temperature dependence coefficient | | | [26] |
| temperature dependence coefficient | | | [26] |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| coefficient for nutrient uptake | | | Defaulted |
| density of total input soluble phosphorus | | | Defaulted |
| density of total input solid phosphorus | | | Defaulted |
| density of solid phosphorus in bottom sediments | | | Defaulted |
| dilution rate | | | Defaulted |
| density limiting coefficient | | | Defaulted |