Skip to main content
Log in

A generalised dynamic model of leaf-level C3 photosynthesis combining light and dark reactions with stomatal behaviour

  • Original Article
  • Published:
Photosynthesis Research Aims and scope Submit manuscript

Abstract

Global food demand is rising, impelling us to develop strategies for improving the efficiency of photosynthesis. Classical photosynthesis models based on steady-state assumptions are inherently unsuitable for assessing biochemical and stomatal responses to rapid variations in environmental drivers. To identify strategies to increase photosynthetic efficiency, we need models that account for the timing of CO2 assimilation responses to dynamic environmental stimuli. Herein, I present a dynamic process-based photosynthetic model for C3 leaves. The model incorporates both light and dark reactions, coupled with a hydro-mechanical model of stomatal behaviour. The model achieved a stable and realistic rate of light-saturated CO2 assimilation and stomatal conductance. Additionally, it replicated complete typical assimilatory response curves (stepwise change in CO2 and light intensity at different oxygen levels) featuring both short lag times and full photosynthetic acclimation. The model also successfully replicated transient responses to changes in light intensity (light flecks), CO2 concentration, and atmospheric oxygen concentration. This dynamic model is suitable for detailed ecophysiological studies and has potential for superseding the long-dominant steady-state approach to photosynthesis modelling. The model runs as a stand-alone workbook in Microsoft® Excel® and is freely available to download along with a video tutorial.

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
Fig. 7

Similar content being viewed by others

References

Download references

Acknowledgements

I am deeply grateful to the Editor of this special issue, Nerea Ubierna Lopez, for editing that improved the clarity and readability, to Joe Quirk for a substantial contribution to writing the first version, I thank Ross Deans (Australian National University, ANU) for unpublished spinach leaf gas exchange data, and Florian Busch (ANU) for help, review, and critical discussion. I am funded through a H2020 Marie Skłodowska-Curie individual fellowship (DILIPHO, ID: 702755).

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Chandra Bellasio.

Ethics declarations

Conflict of interest

I have no conflict of interest.

Electronic supplementary material

Below is the link to the electronic supplementary material.

Supplementary material 1 (PDF 186 KB)

Supplementary material 2 (XLSM 350 KB)

Appendix: model details

Appendix: model details

Flows

A submodel for light reactions of photosynthetic CO2 assimilation in C3 leaves: potential ATP and NADPH production rate

The submodel calculates I1, I2, J1, J2, JATP and JNADPH when fCyc, fPseudocyc, fQ, fNDH, Y(II)LL, s, h, αV, V0V and θV are known. I1 and I2 are the light absorbed by PSI and PSII, respectively. J1 and J2 are the electron flow though PSI and PSII, respectively. JATP and JNADPH are the steady-state rates of ATP and NADPH production, respectively. fCyc is the proportion of electron flow at PSI which follows CEF, fPseudocyc is the fraction of J1 used by alternative electron sinks (APX cycle and nitrate reduction), fQ is the level of Q-cycle engagement, and fNDH is the fraction of fCyc flowing through NDH. The Y(II)LL is the initial yield of PSII extrapolated under zero PPFD, s is the combined energy partitioning coefficient described in Yin et al. (2009), and h is the number of protons required to synthesise each ATP. αV, V0V and θV, define the slope, the offset and the curvature of the function f′(PPFD), expressing the PPFD dependence of Y(II).

When fCyc = 0, I1, I2, J1 and J2 take the values I1,0, I2,0, J1,0 and J2,0, respectively, and J1,0 = J2,0. Then I2,0 and I1,0 can be expressed as (Yin et al. 2004, 2009):

$${I_{2,0}}~={\text{PPFD}} \times s,$$
(1)
$${I_{1,0}}=\frac{{{I_{2,0}}~Y{{({\text{II}})}_{{\text{LL}}}}}}{{Y{{({\text{I}})}_{{\text{LL}}}}}}.$$
(2)

The total light absorbed by both PSI and PSII is I = I1,0 + I2,0 and I < PPFD.

When CEF is engaged, I1 increases by a quantity χ (Yin et al. 2004):

$${I_{1~}}=\left( {1+\chi } \right)~{I_{1,0}},$$
(3)

where χ is calculated as a function of fCyc as (Yin et al. 2004):

$$\chi =\frac{{{f_{{\text{Cyc}}}}}}{{~1+{f_{{\text{Cyc}}}}+~Y{{({\text{II}})}_{{\text{LL}}}}}}.$$
(4a)

I simulate the rate of cyclic electron flow through a tentative function as:

$${f_{{\text{Cyc}}}}=\hbox{max} \left( {0,~ - 1+{{15}^{~~\frac{{v{\text{ATP}}~}}{{{J_{{\text{ATP}}}}}} - ~\frac{{v{\text{NADPH}}}}{{{J_{{\text{NADPH}}}}}}}}} \right).$$
(4b)

When the ratio of actual ATP production relative to the potential \(\left( {\frac{{{v_{{\text{ATP}}}}~}}{{{J_{{\text{ATP}}}}}}} \right)\) is greater than the ratio of actual NADPH production relative to the potential \(\left( {\frac{{{v_{{\text{NADPH}}}}}}{{{J_{{\text{NADPH}}}}}}} \right)\), indicating ATP demand greater than NADPH demand, fCyc will be greater than zero. Equation 4b yields values very close to zero in C3 plants, and further testing will be necessary before application with other photosynthetic types.

If I is constant, I2, J2 and J1 are calculated as (Yin et al. 2004):

$${I_{2~}}=\left( {\frac{1}{{~Y{{({\text{II}})}_{{\text{LL}}}}}} - \chi } \right){I_{2,0}}~~Y{({\text{II}})_{{\text{LL}}}},$$
(5)
$${J_{2~}}={I_{2~}}~Y({\text{II}}),$$
(6)
$${J_{1~}}=\frac{{{J_{2~}}}}{{1 - {f_{{\text{Cyc}}}}}},$$
(7)

where Y(II) is the yield of photosystem II, which depends on PPFD and feedbacks from dark reactions through the novel process-based function:

$$Y({\text{II}})=Y{({\text{II}})_{{\text{LL}}}}~\frac{{v{\text{ATP}}}}{{{J_{{\text{ATP}}}}}}~\frac{{v{\text{NADPH}}}}{{{J_{{\text{NADPH}}}}}}~\left( {1 - \hbox{max} \left[ {0,~f^{\prime}\left( {{\text{PPFD}}} \right)} \right]} \right).$$
(8)

The rationale of Eq. 8 is that Y(II) has a maximum operational value, Y(II)LL, and is quenched by three distinct factors (Müller et al. 2001): (1) the slowing down of ATP synthesis caused by limiting availability of phosphate or ADP (described by \(\frac{{{v_{{\text{ATP}}}}}}{{{J_{{\text{ATP}}}}}}\)); (2) the reduction of the plastoquinone pool (described by \(\frac{{{v_{{\text{NADPH}}}}}}{{{J_{{\text{NADPH}}}}}}\)); (3) reaching the maximum capacity for electron transport (described by f′(PPFD), which responds to PPFD as a non-rectangular hyperbola). f′(PPFD) is calculated with Eq. 20, below, but with different parameterisation, see Table S2. Parameter values were adjusted by fitting modelled assimilation against A/PPFD response curves (Fig. 3, Table S2).

The proton flow to the lumen includes: one proton per electron from water oxidation, one proton from electron flow through the cytochromes (JCyt), two protons from the electron flow through the Q-cycle (JQ, Yin et al. 2004) and two protons from the electron flow through NDH (JNDH, Kramer and Evans 2011). The rate of ATP production is:

$${J_{{\text{ATP~}}}}=\frac{{{J_2}~+~{J_{{\text{Cyt}}}}+~2{J_{Q~}}+~2{J_{{\text{NDH}}}}}}{h},$$
(9)

where h is the number of protons required to synthesise each ATP molecule, the flow through the Q-cycle is JQ = fQJ1; the complement, directly flowing to the b6f complex, is JCyt = (1 − fQ) J1; and the flow through the NDH complex is JNDH = fCycfNDHJ1.

The total NADPH production can be expressed as (Yin et al. 2004):

$${J_{{\text{NADPH}}}}={J_{1~}}\frac{{1 - {f_{{\text{Cyc}}}} - {f_{{\text{Pseudocyc}}}}}}{2},$$
(10a)

The alternative electron sinks include nitrogen metabolism (chiefly reduction) as well as the APX cycle. The APX cycle is known to depend on O2 concentration and the availability of PSI acceptors (Miyake and Yokota 2000; Schreiber and Neubauer 1990). I describe fPseudocyc as a linear function of O2 concentration and \(\frac{{{v_{{\text{NADPH}}}}}}{{{J_{{\text{NADPH}}}}}}\) as:

$${f_{{\text{Pseudocyc}}}}={f_{{\text{Pseudocyc~NR}}}}+4~\left[ {{{\text{O}}_2}} \right]~\left( {1 - \frac{{v{\text{NADPH}}}}{{{J_{{\text{NADPH}}}}}}} \right),$$
(10b)

where the coefficient 4 was fitted empirically to yield a value of fPseudocyc ≈ 0.1 under ordinary ambient conditions, and the fraction of fPseudocyc partitioned to nitrate reduction (fPseudocyc NR) was set at ≈ 0 for simplicity.

The time dependence of JATP and JNADPH was modelled after Bellasio et al. (2017) as:

$$\left\{ {\begin{array}{*{20}{l}} {{J_{{\text{ATP~or~NADPH~}}~t+{\text{d}}t}}={J_{{\text{ATP~or~NADPH}}~t}}~+~\frac{{{J_{{\text{ATP~or~NADPH}}~}} - {J_{{\text{ATP~or~NADPH}}~t}}}}{{{K_{J{~_{{\text{ATP~or~NADPH}}}}}}}}~~{\text{d}}t}&{~{\text{if}}~{J_{{\text{ATP~or~NADPH}}~t}}<{J_{{\text{ATP~or~NADPH}}}}} \\ {{J_{{\text{ATP~or~NADPH }}t}}={J_{{\text{ATP~or~NADPH}}}}}&{{\text{else}}} \end{array}} \right.,$$
(11)

where KJ ATP or NADPH is the time constant for an increase in JATP or JNADPH, JATP or NADPH t+dt and JATP or NADPH t are the values at the time step t + dt or at the previous step t, respectively; JATP or NADPH are the steady-state values (Eqs. 9 and 10a).

Actual rates of ATP and NADPH production

The actual rates of ATP (vATP) and NADPH (vNADPH) production are calculated after Wang et al. (2014b):

$${v_{{\text{ATP}}}}=\frac{{{J_{{\text{ATP}}}}~\left( {\left[ {{\text{ATP}}} \right]\left[ {{P_{\text{i}}}} \right] - \frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{e}}~}}}}} \right)}}{{{K_{{\text{m~ADP}}}}{K_{{\text{m~}}{P_{\text{i}}}}}\left( {1+\frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{m~ADP}}}}}}+\frac{{\left[ {{\text{ATP}}} \right]}}{{{K_{{\text{m~ATP}}}}}}+\frac{{\left[ {{P_{\text{i}}}} \right]}}{{{K_{{\text{m~}}{P_{\text{i}}}}}}}+\frac{{\left[ {{\text{ATP}}} \right]\left[ {{P_{\text{i}}}} \right]}}{{{K_{{\text{m~ADP}}}}{K_{{\text{m~}}{{\text{P}}_{\text{i}}}}}}}} \right)}},$$
(12)
$${v_{{\text{NADPH}}}}=\frac{{{J_{{\text{NADPH}}}}~\left( {\left[ {{\text{NADP}}} \right] - \frac{{\left[ {{\text{NADPH}}} \right]}}{{{K_{{\text{e}}~}}}}} \right)}}{{{K_{{\text{m~NADP}}}}\left( {1+\frac{{\left[ {{\text{NADP}}} \right]}}{{{K_{{\text{m~NADP}}}}}}+\frac{{\left[ {{\text{NADPH}}} \right]}}{{{K_{{\text{m~NADPH}}}}}}} \right)}},$$
(13)

where square brackets indicate metabolite concentration, Km represents the Michaelis–Menten constant for a given metabolite, and Ke is the equilibrium constant of the reaction (Table S1).

CO2 diffusion, dissolution and hydration

The rate of CO2 diffusion through the stomata is:

$${\text{C}}{{\text{O}}_2}~{\text{stomatal~diffusion}}=\frac{{{g_{\text{S}}}~\left( {{C_{\text{a}}} - {C_{\text{i}}}} \right)}}{{1000}},$$
(14)

where gS is stomatal conductance to CO2 (mol m−2 s−1); Ca and Ci are the CO2 concentrations (µmol mol−1) external to the leaf and in the intercellular space, respectively; and the 1000 is used to convert the units from micromoles to mmol m−2 s−1.

The rate of CO2 dissolution in aqueous media within the leaf is:

$${\text{C}}{{\text{O}}_2}~{\text{dissolution}}=\frac{{{g_{\text{M}}}~\left( {{C_{\text{i}}} - \left[ {{\text{C}}{{\text{O}}_2}} \right]{K_{{\text{hC}}{{\text{O}}_2}~}}} \right)}}{{1000}},$$
(15)

where gM is mesophyll conductance to CO2 diffusion (mol m−2 s−1), [CO2] is the CO2 concentration in mesophyll cells (mM), which is assumed to be spatially uniform, \({K_{{\text{hC}}{{\text{O}}_2}}}\) is CO2 volatility (the reciprocal of solubility) (µbar mM−1) and 1000 is used to convert the units into mmol m−2 s−1, the unit of all subsequent rates.

The rate of CO2 hydration to bicarbonate is (Wang et al. 2014b):

$${\text{CA}}=\frac{{{V_{{\text{MAX~CA}}}}~\left( {\left[ {{\text{C}}{{\text{O}}_2}} \right] - \frac{{\left[ {{\text{HC}}{{\text{O}}_3}^{ - }} \right]\left[ {{{\text{H}}^+}} \right]}}{{{K_{{\text{e~}}}}}}} \right)}}{{{K_{{\text{m~C}}{{\text{O}}_2}}}\left( {1+\frac{{\left[ {{\text{C}}{{\text{O}}_2}} \right]}}{{{K_{{\text{m~C}}{{\text{O}}_2}}}}}+\frac{{\left[ {{\text{HC}}{{\text{O}}_3}^{ - }} \right]}}{{{K_{{\text{m~HC}}{{\text{O}}_3}}}}}} \right)}},$$
(16)

where VMAX CA is the maximum hydration rate.

Reaction rates

The rate of Rubisco carboxylation (VC) was modified from Wang et al. (2014b) as:

$${V_{\text{C}}}=\frac{{{V_{{\text{C~MAX}}}}~~{R_{{\text{act}}}}~~f\left( {{\text{RuBP}}} \right)~\left[ {{\text{RuBP}}} \right]\left[ {{\text{C}}{{\text{O}}_2}} \right]}}{{\left( {K_{{{\text{m~C}}{{\text{O}}_2}}}^{'}+\left[ {{\text{C}}{{\text{O}}_2}} \right]} \right)\left( {K_{{{\text{m~RuBP}}}}^{'}+\left[ {{\text{RuBP}}} \right]} \right)}},$$
(17)

where VC MAX is the maximum carboxylation rate. In the VC MAX used in Farquhar et al. (1980), Rubisco is assumed fully activated and also fully RuBP saturated in the ‘enzyme-limited’ case. Here, VC MAX is more closely comparable to the in vitro rate. Ract is the Rubisco activation state, a time-dependent variable calculated as:

$${R_{{\text{act~}}t{\text{+d}}t}}={R_{{\text{act}}~t}}+\left\{ {\begin{array}{*{20}{l}} {\frac{{{R_{{\text{act~eq}}}} - ~{R_{{\text{act}}~t}}}}{{{\tau _{\text{i}}}}}~{\text{d}}t}&{i{\text{f}}~{R_{{\text{act}}~t}}<{R_{{\text{act~eq}}}}} \\ {\frac{{{R_{{\text{act~eq}}}} - ~{R_{{\text{act~}}t}}}}{{{\tau _{\text{d}}}}}~{\text{d}}t}&{{\text{else}}} \end{array}} \right.,$$
(18)

where τi and τd are the time constants for Rubisco induction and deactivation, respectively (Seemann et al. 1988), and the steady state Ract value is:

$${R_{{\text{act~}}}}=f\left( {{\text{PPFD}}} \right)~~~f\left( {\left[ {{\text{C}}{{\text{O}}_2}} \right]} \right),$$
(19)

where \(f\left( {{\text{PPFD}}} \right)\) simulates activation state of Rubisco independently of CO2 concentration, and I included f([CO2]) to capture the inactivation of Rubisco observed in vivo at low CO2. The \(f\left( {{\text{PPFD}}} \right)\) and f([CO2]) were modelled with non-rectangular hyperbolas (Gross et al. 1991):

$$f\left( {{\text{PPFD}}} \right)~=~{V_0}+\frac{{{\alpha _{\text{V}}}~{\text{PPFD}}+1 - {V_0}~ - \sqrt {{{\left( {{\alpha _{\text{V}}}~{\text{PPFD}}+1 - {V_0}} \right)}^2} - 4{\alpha _{\text{V}}}~{\text{PPFD}}~{\theta _{\text{V}}}} }}{{2~{\theta _{\text{V}}}}},$$
(20)
$$f\left( {\left[ {{\text{C}}{{\text{O}}_2}} \right]} \right)~=~{V_{0{\text{C}}}}+\frac{{{\alpha _{\text{C}}}~\left[ {{\text{C}}{{\text{O}}_2}} \right]+1 - {V_{0{\text{C}}}}~ - \sqrt {{{\left( {{\alpha _{\text{V}}}~\left[ {{\text{C}}{{\text{O}}_2}} \right]+1 - {V_{0{\text{C}}}}} \right)}^2} - 4{\alpha _{\text{C}}}~\left[ {{\text{C}}{{\text{O}}_2}} \right]~{\theta _{\text{C}}}} }}{{2~{\theta _{\text{C}}}}},$$
(21)

where V0, αV, and θV are empirical parameters of the hyperbola for f(PPFD) defining the initial activity in the dark, the slope of the dependency and the curvature, respectively; V0C, αC, and θC are the equivalent parameters for f([CO2]).

The f(RuBP) is a function of RuBP concentration, relative to the concentration of Rubisco active sites, which was modelled using a non-rectangular hyperbola after Farquhar et al. (1980):

$$f\left( {{\text{RuBP}}} \right)~=~\frac{{{E_{\text{T}}}~+~K_{{{\text{m RuBP}}}}^{'}+\left[ {{\text{RuBP}}} \right] - \sqrt {{{\left( {{E_{\text{T}}}~+~K_{{{\text{m RuBP}}}}^{'}+\left[ {{\text{RuBP}}} \right]} \right)}^2} - 4\left[ {{\text{RuBP}}} \right]{E_{\text{T}}}} }}{{2{E_{\text{T}}}}}~,$$
(22)

where ET is the total concentration of Rubisco catalytic sites, calculated from VC MAX and turnover rate after Wang et al. (2014b). The Michaelis–Menten constant for RuBP and CO2 are:

$$K_{{{\text{m RuBP}}}}^{\prime }={K_{{\text{m RuBP}}}}\left( {1+\frac{{\left[ {{\text{PGA}}} \right]}}{{{K_{{\text{m~PGA}}}}}}+\frac{{\left[ {{\text{NADP}}} \right]}}{{{K_{{\text{i~NADP}}}}}}+\frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{i~ADP}}~}}}}+\frac{{\left[ {{P_{\text{i}}}} \right]}}{{{K_{{\text{i~}}{P_{\text{i}}}{\text{~}}}}}}} \right),$$
(23)
$$K_{{{\text{m~C}}{{\text{O}}_2}}}^{\prime }={K_{{\text{m~C}}{{\text{O}}_2}}}\left( {1+\frac{{\left[ {{{\text{O}}_2}} \right]}}{{{K_{{\text{m~}}{{\text{O}}_2}}}}}} \right),$$
(24)

where Ki are the constants for the competitive inhibition.

The rate of Rubisco oxygenation (VO) was calculated after Farquhar et al. (1980) as:

$${V_{\text{O}}}={V_{\text{C}}}~2{\gamma ^*}~\frac{{\left[ {{{\text{O}}_2}} \right]}}{{\left[ {C{{\text{O}}_2}} \right]}},$$
(25)

where γ* is half the reciprocal Rubisco specificity, calculated in the liquid phase (von Caemmerer 2000) using constants from Sander (2015) and Warneck and Williams (2012). In this model, the glycine decarboxylase (GDC) decarboxylation rate equals VO; for a justification and possible stoichiometric variants, see Bellasio (2017).

The rate of RuP phosphorylation was modified from Wang et al. (2014b) as:

$${\text{Ru}}{{\text{P}}_{{\text{Phosp}}}}=\frac{{{V_{{\text{MAX~}}}}~\left[ {{\text{ATP}}} \right]\left[ {{\text{RuP}}} \right]~ - \frac{{\left[ {{\text{ATP}}} \right]\left[ {{\text{RuBP}}} \right]}}{{{K_{{\text{e~}}}}}}}}{{\left( {\left[ {{\text{ATP}}} \right]~+~{K_{{\text{m~ATP}}}}\left( {1+~\frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{i~ADP}}}}}}} \right)} \right)\left( {\left[ {{\text{RuP}}} \right]+{K_{{\text{m~RuP}}}}\left( {1+~\frac{{\left[ {{\text{PGA}}} \right]}}{{{K_{{\text{i~PGA}}}}}}~+~\frac{{\left[ {{\text{RuBP}}} \right]}}{{{K_{{\text{i~RuBP}}}}}}~+~\frac{{\left[ {{P_{\text{i}}}} \right]}}{{{K_{{\text{i~}}{P_{\text{i}}}}}}}} \right)} \right)}}.$$
(26)

The reducing phase of the reductive pentose phosphate pathway was modelled as a single-step pseudoreaction. The rate of PGA reduction (PR) was calculated by fusing the rates of PGA phosphorylation and DPGA reduction from Wang et al. (2014b) as:

$${\text{PR}}=\frac{{{V_{{\text{MAX~}}}}~\left[ {{\text{ATP}}} \right]\left[ {{\text{PGA}}} \right]\left[ {{\text{NADPH}}} \right]~}}{{\left( {\left[ {{\text{PGA}}} \right]~+~{K_{{\text{m~PGA}}}}\left( {1+~\frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{i~ADP}}}}}}} \right)} \right)\left( {\left[ {{\text{ATP}}} \right]~+~{K_{{\text{m~ATP}}}}\left( {1+~\frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{i~ADP}}}}}}} \right)} \right)\left( {\left[ {{\text{NADPH}}} \right]~+~{K_{{\text{m~NADPH}}}}\left( {1+~\frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{i~ADP}}}}}}} \right)} \right)}}.$$
(27)

The constants in Eq. 27 were adapted from the original separate reactions in Wang et al. (2014b) to maintain physiologically realistic concentrations of product and substrates. This was necessary to account for neglecting phosphorylated intermediates of the RPP.

Carbohydrate synthesis was assumed to be a single-step reaction, and the rate was calculated by simplifying the combined processes of starch and sucrose synthesis from Wang et al. (2014b) as:

$${\text{CS}}=\frac{{{V_{{\text{MAX~}}}}~\left( {\left[ {{\text{DHAP}}} \right] - 0.4} \right)\left( {1 - \frac{{\left| {{\text{PR}}} \right|~\left[ {{P_{\text{i}}}} \right]}}{{{K_{{\text{e~}}}}}}} \right)}}{{\left( {\left[ {{\text{DHAP}}} \right]~+~{K_{{\text{m~DHAP}}}}\left( {1+~\frac{{\left[ {{\text{ADP}}} \right]}}{{{K_{{\text{i~ADP}}}}}}} \right)} \right)}}.$$
(28a)

With Eq. 28a carbohydrates are synthesised when [DHAP] > 0.4 mM following a saturating Michaelis–Menten kinetics, inhibited by ADP. Concentration of sucrose, starch and their precursor are not calculated. To capture the reversible nature of the original Wang et al. (2014b) formulation, I use the quantity \(\frac{{\left| {{\text{PR}}} \right|~\left[ {{P_{\text{i}}}} \right]}}{{{K_{{\text{e~}}}}}}\), where \(\left| {{\text{PR}}} \right|\) ‘senses’ the concentration of sucrose and starch, hypothesised to be proportional to the rate of DHAP synthesis.

The interconversion phase of the RPP was modelled as a single-step pseudoreaction through a generic Michaelis Menten equation for equilibrium reaction (Zhu et al. 2007) as:

$${\text{RPP=}}\frac{{{V_{{\text{MAX~}}}}{\text{~}}\left[ {{\text{DHAP}}} \right]{\text{~}}\left( {{\text{1}} - \frac{{\left| {{\text{RuP}}} \right|}}{{{K_{{\text{e~}}}}}}} \right)}}{{\left( {\left[ {{\text{DHAP}}} \right]{\text{~+~}}{K_{{\text{m~DHAP}}}}} \right)}}{\text{.}}$$
(28b)

The constants in Eq. 28b were adapted from the original separate reactions in Wang et al. (2014b) to operate and maintain physiologically realistic concentrations of substrates.

Stocks

Change in metabolite concentrations

The change in metabolite concentrations in time \(\frac{{{\text{d}}\left[ {} \right]}}{{{\text{d}}t}}~\) was described by a set of ordinary differential equations based on the stoichiometry of Bellasio (2017) informed with the reaction rates described above and converted from variation in leaf-level pool to variation in concentration using the mesophyll volume as described in Wang et al. (2014b).

The rates of change in concentrations of CO2, bicarbonate (HCO3), RuBP, PGA, DHAP, ATP and NADPH were calculated as:

$$\frac{{{\text{d}}\left[ {{\text{C}}{{\text{O}}_2}} \right]}}{{{\text{d}}t}}{V_{\text{M}}}={\text{C}}{{\text{O}}_2}~{\text{Dissolution}}+{R_{{\text{LIGHT}}}} - {V_{\text{C}}}+0.5~{\text{GDC}} - {\text{CA,}}$$
(29)
$$\frac{{{\text{d}}\left[ {{\text{HC}}{{\text{O}}_3}^{ - }} \right]}}{{{\text{d}}t}}{V_{\text{M}}}={\text{CA,}}$$
(30)
$$\frac{{{\text{d}}\left[ {{\text{RuBP}}} \right]}}{{{\text{d}}t}}{V_{\text{M}}}={\text{Ru}}{{\text{P}}_{{\text{Phosp}}}} - ~{V_{\text{C}}} - {V_{\text{O}}},$$
(31)
$$\frac{{{\text{d}}\left[ {{\text{PGA}}} \right]}}{{{\text{d}}t}}{V_{\text{M}}}=2~{V_{\text{C}}}+{V_{\text{O}}}+0.5{\text{~GDC}} - {\text{PR}} - \frac{1}{3}{R_{{\text{LIGHT}}}},$$
(32)
$$\frac{{{\text{d}}\left[ {{\text{DHAP}}} \right]}}{{{\text{d}}t}}{V_{\text{M}}}={\text{PR}} - {\text{CS}} - \frac{5}{3}{\text{Ru}}{{\text{P}}_{{\text{Phosp}}}},$$
(33)
$$\frac{{{\text{d}}\left[ {{\text{ATP}}} \right]}}{{{\text{d}}t}}{V_{\text{M}}}={v_{{\text{ATP}}}} - {\text{Ru}}{{\text{P}}_{{\text{Phosp}}}} - {V_{\text{O}}} - {\text{~PR}} - 0.5~{\text{CS,}}$$
(34)
$$\frac{{{\text{d}}\left[ {{\text{NADPH}}} \right]}}{{{\text{d}}t}}{V_{\text{M}}}={v_{{\text{NADPH}}}} - ~{\text{PR}} - 0.5~{V_{\text{O}}},$$
(35)
$$\frac{{{\text{d}}\left[ {{\text{RuP}}} \right]}}{{{\text{d}}t}}{V_{\text{M}}}={\text{RPP}} - ~{\text{Ru}}{{\text{P}}_{{\text{Phosp}}}},$$
(36)

where VM is mesophyll volume per meter square of leaf (L m−2) calculated after considering the leaf half-full of mesophyll (Lawlor 1993), RLIGHT is light respiration and is input to the model as described in Bellasio (2017), and all the other flux rates have been previously described: CO2 dissolution (Eq. 15), CA (Eq. 16), RuPPhosph (Eq. 26), VC (Eq. 17), VO (= GDC, Eq. 25), PR (Eq. 27), CS (Eq. 28a), vATP (Eq. 12), vNADPH (Eq. 13) and RPP (Eq. 28b). Equations 2930 were derived in this study and Eqs. 3136 are modified from (Bellasio 2017).

Concentrations determined from total metabolite pools

The concentrations of ADP, NADP+ and phosphate (\(\left[ {{P_{\text{i}}}} \right]\)) are calculated simply by subtraction from a total pool:

$$\left[ {{\text{ADP}}} \right]={A_{{\text{Tot}}}} - \left[ {{\text{ATP}}} \right],$$
(37)
$$\left[ {{\text{NAD}}{{\text{P}}^+}} \right]={N_{{\text{Tot}}}} - \left[ {{\text{NADPH}}} \right],$$
(38)
$$\left[ {{P_{\text{i}}}} \right]={P_{{\text{iTot}}}} - \left[ {{\text{PGA}}} \right]~ - \left[ {{\text{DHAP}}} \right] - \left[ {{\text{RuP}}} \right] - 2\left[ {{\text{RuBP}}} \right]~ - \left[ {{\text{ATP}}} \right],$$
(39)

where ATot, NTot, and PiTot are the total pools of adenylates, nicotinamides and phosphate, respectively.

The hydro-mechanical model of stomatal behaviour

The model calculates gS after Bellasio et al. (2017) as:

$${g_{\text{S}}}=\hbox{max} \left( {{g_{{\text{S}}~0}},~\frac{{\chi ~\beta ~\tau ~\left( {{\Psi _{{\text{Soil}}}}+{\pi _{\text{e}}}} \right)}}{{1+\chi ~\beta ~\tau ~{R_{\text{h}}}~{D_{\text{S}}}}}} \right),$$
(40)

where χβ is a combined parameter scaling turgor-to-conductance and the hydro-mechanical-to-biochemical response; τ is the sensor of biochemical forcing; ΨSoil is soil water potential; πe is epidermal osmotic pressure; Rh is the effective hydraulic resistance to the epidermis, calculated as 1/Kh, the corresponding hydraulic conductance; and DS is the leaf-to-boundary layer H2O mole fraction gradient, a measure of vapour pressure deficit, VPD. The parameter τ encompasses the biochemical components of the model and is calculated from f(RuBP) as:

$$\tau ={\tau _0}+f\left( {{\text{RuBP}}} \right),$$
(41)

where τ0, the basal level of τ, was manually assigned. Stomatal dynamics were accounted for by describing the time dependence of gS with a set of recursive equations (Bellasio et al. 2017):

$${g_{{\text{S}}~t+{\text{dt}}}}={g_{{\text{S}}~t}}+\left\{ {\begin{array}{*{20}{l}} {\frac{{{g_{{\text{S}}~}} - {g_{{\text{S}}~t}}}}{{{K_{\text{i}}}}}~{\text{d}}t}&{i{\text{f}}~{g_{{\text{S}}~t}}<{g_{\text{S}}}} \\ {\frac{{{g_{{\text{S}}~}} - {g_{{\text{S}}~t}}}}{{{K_{\text{d}}}}}~{\text{d}}t~}&{{\text{else}}} \end{array}} \right.,$$
(42)

where gSt+dt and gSt are the gS values at the time step t + dt or at the previous step t, respectively; gS is the steady-state value (Eq. 40), Ki and Kd are the time constants for an increase and decrease in gS, respectively.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bellasio, C. A generalised dynamic model of leaf-level C3 photosynthesis combining light and dark reactions with stomatal behaviour. Photosynth Res 141, 99–118 (2019). https://doi.org/10.1007/s11120-018-0601-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11120-018-0601-1

Keywords

Navigation