Articles | Volume 20, issue 13
https://doi.org/10.5194/bg-20-2671-2023
https://doi.org/10.5194/bg-20-2671-2023
Research article
 | 
06 Jul 2023
Research article |  | 06 Jul 2023

A differentiable, physics-informed ecosystem modeling and learning framework for large-scale inverse problems: demonstration with photosynthesis simulations

Doaa Aboelyazeed, Chonggang Xu, Forrest M. Hoffman, Jiangtao Liu, Alex W. Jones, Chris Rackauckas, Kathryn Lawson, and Chaopeng Shen
Abstract

Photosynthesis plays an important role in carbon, nitrogen, and water cycles. Ecosystem models for photosynthesis are characterized by many parameters that are obtained from limited in situ measurements and applied to the same plant types. Previous site-by-site calibration approaches could not leverage big data and faced issues like overfitting or parameter non-uniqueness. Here we developed an end-to-end programmatically differentiable (meaning gradients of outputs to variables used in the model can be obtained efficiently and accurately) version of the photosynthesis process representation within the Functionally Assembled Terrestrial Ecosystem Simulator (FATES) model. As a genre of physics-informed machine learning (ML), differentiable models couple physics-based formulations to neural networks (NNs) that learn parameterizations (and potentially processes) from observations, here photosynthesis rates. We first demonstrated that the framework was able to correctly recover multiple assumed parameter values concurrently using synthetic training data. Then, using a real-world dataset consisting of many different plant functional types (PFTs), we learned parameters that performed substantially better and greatly reduced biases compared to literature values. Further, the framework allowed us to gain insights at a large scale. Our results showed that the carboxylation rate at 25 C (Vc,max25) was more impactful than a factor representing water limitation, although tuning both was helpful in addressing biases with the default values. This framework could potentially enable substantial improvement in our capability to learn parameters and reduce biases for ecosystem modeling at large scales.

Dates
1 Introduction

Plant photosynthesis is critically important for regulating the global carbon and nutrient cycles, and thus the future climate. Understanding future climate trajectories requires understanding photosynthetic responses to changes in environmental factors including atmospheric CO2 concentrations, radiation, temperature, humidity, and nutrient and water availability (Kirschbaum, 2004). Photosynthesis is influenced by many factors such as higher CO2 levels, reduced productivity of vegetation (i.e., nutrient concentration; Thompson et al., 2017), intensified droughts (Urban et al., 2017; Xu et al., 2019), and rising temperatures (Dusenge et al., 2019) under a changing climate. To comprehensively evaluate the impacts of these changing processes and vegetation feedback to the atmosphere, we need accurate representations of photosynthesis in our models.

For global assessments of the carbon cycle, vegetation models were developed to simulate terrestrial ecosystem processes and the distributions of vegetation, both vertically in the soil–plant system and horizontally across the landscape. Substantial efforts over the last few decades have improved the representation of vegetation and its responses and feedback to climate change (Fisher et al., 2018). A typical framework structure for vegetation models is to keep track of changes in carbon and optionally nutrient states, driven by climatic variables and modulated by soil properties, with feedback to the climate, e.g., CO2 releases, radiation, and vegetation composition and structure. A core component of the vegetation module is photosynthesis (Quillet et al., 2010).

Present ecosystem models for photosynthesis are based primarily on mechanistic descriptions of plant photosynthesis pathways, but this theoretically sound modeling paradigm faces many challenges, with parametric uncertainty being a major one. Photosynthesis models may describe limitations of carboxylation rates, light availability, and plant-specific factors like enzyme efficiencies for C3 and C4 plants differently (Farquhar et al., 1980; Farquhar and Caemmerer, 1982; Meyer, 1983; Von Caemmerer, 2003, 2013; Yin and Struik, 2009). They contain many parameters that quantify these efficiencies and limitations. In the past, these parameters have been estimated from different approaches: (1) obtained from a limited set of in situ sites and scaled based on climate and environmental factors (Verheijen et al., 2013), (2) calibrated on observational data site by site or for a few sites for a specific plant functional type (PFT; Mäkelä et al., 2019; Wang et al., 2014), or (3) optimized based on environmental conditions (Ali et al., 2016). However, these estimated values may not be optimal at the global scale. Site-by-site calibrations using genetic or similar algorithms are highly expensive and are limited in their spatial coverage and generalizability to different PFTs and species. Furthermore, such calibration faces the issue of nonuniqueness (which some call equifinality; Beven and Freer, 2001), where different parameter sets produce the same outcome. As a result, calibration can easily lead to poorly generalizable parameter values. This problem exists for many domains with diverse parameters, including ecosystem modeling (Tang and Zhuang, 2008). It has been similarly found in hydrologic modeling and has troubled scientists there for decades (Beven, 2006). More recently, it has been possible for some parameters to be fitted directly from large datasets with directly measured parameter values (Luo et al., 2021), which is highly valuable but is limited to those parameters with extensive observations, e.g., soil water retention and hydraulic properties. An efficient way to permit large-scale inverse modeling is needed.

There has been substantial progress in utilizing modern machine learning (ML) for geosciences. Purely data-driven deep learning models (LeCun et al., 2015; Reichstein et al., 2019; Shen, 2018; Shen et al., 2018) directly learn from data so they tend to be fairly accurate, and many have outperformed traditional models for a large number of applications across domains such as hydrology (Feng et al., 2020, 2021; Rahmani et al., 2020, 2021; Liu et al., 2023; Wunsch et al., 2021; Fang and Shen, 2020), agriculture (ElSaadani et al., 2021; Hossain et al., 2019; Liu et al., 2022; Saleem et al., 2019), energy balance (Zhu et al., 2021), cryosphere (Leong and Horgan, 2020; Zhang et al., 2019), water quality (Hrnjica et al., 2021; Zhi et al., 2021; Saha et al., 2023; Zhi et al., 2023), and ecosystem modeling (Zhang et al., 2020, 2021). In particular, the loss function is defined on all data points and sites, enforcing stronger constraints and leading to the data synergy effect, where the model gains better performance as the amount of training data increases (Fang et al., 2022). Unfortunately, deep learning models also lack interpretability and process clarity, and can only output trained variables with extensive observations. This need for data is often not satisfied for ecological processes.

To aid geoscientific models in general, Tsai et al. (2021) presented an efficient framework known as differentiable parameter learning (dPL), in an effort to leverage recent progress in ML to mitigate the issues with parameter inversion discussed above. This framework turns parameter estimation into a large-scale ML problem. It is mainly composed of a parameter estimation module based on a neural network (NN), combined with a process-based model (PBM) or its surrogate. The whole framework must be “programmatically differentiable” (Baydin et al., 2018; Innes et al., 2019), which refers to a programming paradigm that can efficiently and accurately obtain the gradients of the outputs with respect to any of the variables used in the model (Shen et al., 2023). Once we have programmatic differentiability, dPL can efficiently learn unknown functions from big data to serve as either a parameterization or process representation. Tsai et al. (2021) found that this framework scales well with more data, produces spatially and temporally well-generalized parameter sets, extends well to uncalibrated variables, and saves orders of magnitude in computational time. Feng et al. (2022a) further showed that a differentiable process-based hydrologic model with dPL could approach the performance of a purely data-driven ML model, and potentially outperform ML in data-sparse regions (Feng et al., 2022b). These successes can be conveniently migrated to the ecosystem modeling domain.

Here, we applied the dPL framework to the leaf-scale photosynthesis module of the Functionally Assembled Terrestrial Ecosystem Simulator (FATES) model. FATES is an ecosystem model that describes coexistence and competition in PFTs (Koven et al., 2020). FATES can be used as an ecosystem module in the Community Land Model (CLM; Oleson et al., 2013; Lawrence et al., 2019) to represent ecosystem demography (Fisher et al., 2015). The photosynthesis module is based on the Farquhar photosynthesis model. To apply the dPL framework in our study, we first reimplemented the photosynthesis module from FATES so that it became programmatically differentiable. Second, we connected this model to NNs for parameter estimation. With this tool, we aimed to answer the following questions.

  1. Are parameters like Vc,max25 and soil water limitation factor simultaneously identifiable?

  2. What is the achievable model performance, in terms of predicting photosynthesis rates in space and time, by tuning the parameters for the classical photosynthesis module without changing the model structure?

  3. Are parameters learned from a large global dataset similar to the values used in current models?

In the following, we first describe the photosynthesis model with different parameter estimation experiments and target datasets. We then discuss the parameters chosen to be estimated and their significance. Afterward, we present the results from synthetic experiments and experiments based on real datasets from sites around the globe. Finally, we compare the learned parameters to values from the literature and provide some suggestions for future work.

2 Methods and datasets

2.1 General overview

Our general framework trains connected NNs to provide parameters (and later process representations) to PBMs, in this case the photosynthesis module of the FATES ecosystem model, on all the training data points simultaneously (Fig. 1a). The NNs map from some raw inputs to some tuneable physical parameters (θ; later extensible to processes) required for the PBM. The predicted physical parameters are then fed into the differentiable PBM along with other required forcing variables (F) and untuned constant attributes (θc) to compute the simulated target variable (ysim), which is compared with observations to compute a loss function. The forward run starts from the NN inputs and ends at the loss function (following the blue arrows in Fig. 1a); the model's goal is to minimize the output of the loss function. We then backpropagate the errors (shown by black arrows in Fig. 1a) through the PBM equations back to the NNs so we can train them using gradient descent. To support gradient-based training, the entire framework must be differentiable (Shen et al., 2023), which ensures that neither the NN nor the PBM is a black box – they both allow explicit inspection and modification of the internal structures. Thus, the photosynthesis module of FATES had to be reimplemented on a differentiable platform.

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f01

Figure 1Diagram showing the dPL framework, which is a hybrid of NNs and the photosynthesis module in the FATES ecosystem model written on a differentiable platform. (a) The generic workflow: some raw information is mapped into physical parameters via an NN. These parameters are sent into a PBM, which then outputs the simulations, ysim, that are compared with the observations, yobs. Direct supervision for the physical parameters themselves is not required – we do not need ground truth values for these parameters. The loss function is “global” in that it involves all training data points, rather than being computed site-by-site as done in traditional calibration. (b) The workflow for the computational example described in this work. We estimate either Vc,max25 or the parameter B, or both of them at the same time, using NNs. The parameters are then fed into the differentiable photosynthesis module in FATES, which then outputs the net photosynthesis rate, An(sim), that is compared with An(obs). When not estimated from the data, default values from the literature were used. Blue arrows show the running of the NNs with the PBM in a forward (“prediction”) mode, while black arrows indicate backpropagation from the loss function back through the differentiable model equations to the NNs to update their weights, which is only done during initial NN training.

Download

In this case, the PBM is the photosynthesis module in FATES, which can be written as a nonlinear system of equations, and its solution is implicit. The system can be written as

(1) f x ; θ , θ c , F = 0 ; y = h x , θ , θ c , F ,

where f represents the physical system constraint, h is an observation operator, x represents the unknowns of the equations (in this case the intercellular leaf CO2 partial pressure, Pa), y is an observable variable (in this case net photosynthetic rate, µmol m−2 s−1) that is dependent on x via h, F represents some meteorological forcing variables such as radiation and air temperature, θ represents a list of tuneable physical parameters, and θc represents untuned constant attributes. Given a set of θ with known θc and F, we need to solve for x from f and send the solution into h to further compute y: y=hf-1θ,θc,F,θ,θc,F.

This whole workflow can be lumped into one model:

(2) y = δ psn θ , θ c , F

where δpsn represents the overall photosynthesis model. Some of the tuneable parameters are typically formulated as being PFT dependent (e.g., the maximum carboxylation rate at 25 C, Vc,max25), where each PFT includes groups of plant species that share similar physical and phenological characteristics leading to similar interactions with the environment. Other tuneable parameters are related to soil water availability (e.g., soil water stress parameters). We posit that there exists a parameterization scheme, θ=gW(R), which is a mapping relationship of some underlying attributes R (e.g., soil attributes and plant traits) with the physical parameters represented by NN g with W learnable weights. Thus, we can learn W so that the simulated variable y matches the observations y*:

(3) W = argmin W ( L δ psn g W R , θ c , F , y ) ,

where L is the loss function. For the purpose of solving the inverse problem and training the NN gW in an “online” mode using gradient descent (the only practically employed algorithm for NN training), we reimplemented the photosynthesis module in FATES onto two differentiable platforms: Julia and PyTorch (discussed in more detail below).

In order to test the learning capability of our framework and the identifiability of the parameters, we first ran synthetic experiments to verify if the model would be able to correctly retrieve assumed values for the physical parameters. Second, using a dataset with thousands of photosynthesis rate measurements, we trained the differentiable model to obtain estimated parameters at the global scale, and compared them to the literature.

2.2 The Farquhar photosynthesis model

The FATES photosynthesis module is based on the classical Farquhar model for C3 plants (Farquhar et al., 1980), which calculates the photosynthetic rate based on carbon fluxes under different limitations. For C4 plants, it uses the Collatz model (Collatz et al., 1992). Both models assume that the gross photosynthetic rate is affected by the maximum rate of carboxylation and is limited by the concentration of RuBP carboxylase (Rubisco; Ac), light and electron transport (Aj), and the concentration of PEP carboxylase enzyme in C4 plants (Ap). The final gross photosynthetic rate “A” is calculated using the empirical curvature parameters (θcj and θip), while the net photosynthetic rate An is the same as the gross rate (A) after the plant respiration rate (Rd) is subtracted. The system can be described succinctly as the following, with Eqs. (4) and (5) playing the roles of f and h in Eq. (1), respectively; the whole set of associated equations is detailed in Fig. 2 and Appendix A.

(4)Ci=Ca-AnPatm(1.4gs+1.6gb)(gs×gb),(5)An=ACi-Rd.

Equation (4) is a single-variable nonlinear equation, with the intercellular leaf CO2 pressure (Ci) as the unknown term to be solved (serving as the x term in Eq. 1). Ci is influenced by the CO2 partial pressure near the leaf surface (Ca), the net photosynthetic rate (An), the atmospheric pressure (Patm), the leaf stomatal conductance (gs), and the leaf boundary layer conductance (gb). Upon solving for Ci, we can further calculate An, which is the y term in Eq. (1). In the original implementations of FATES and CLM, the system of nonlinear equations was solved iteratively using fixed-point iteration (Oleson et al., 2013).

In order to train the physical equations and NNs together using gradient descent, the above equations were implemented on differentiable platforms to support backpropagation. We developed two alternative implementations: using PyTorch in Python (Paszke et al., 2019) and using Julia (Bezanson et al., 2012). The PyTorch version solves the coupled nonlinear equations using our parallel implementation of Newton iteration with automatic differentiation, while the Julia version uses adjoint-based methods implemented via a symbolic computer algebra system and is compatible with a wide variety of nonlinear solvers (Gowda et al., 2022). In contrast to the previous fixed-point iteration used by the original FATES model, our PyTorch Newton iteration solver can run on a graphical processing unit (GPU) in parallel for many sites. Newton's iteration method features second-order convergence compared to the slower convergence of fixed-point iteration, while GPU parallelism represents orders-of-magnitude savings in computational time compared to the original algorithm in FATES. The photosynthesis problem studied here has only one unknown (Ci) even though there are many other supporting equations, but we have successfully tested other larger nonlinear systems. Altogether, we can train this model with the coupled NNs for hundreds of data points in under 10 min (typically in 600 iterations) and could also train the model on 10 000 data points. For future work where time steps are involved, the adjoint method will likely be employed to reduce GPU memory use during nonlinear iterative solving. For the Julia implementation, the symbolic toolbox ModelingToolKit.jl (Gowda et al., 2022; Ma et al., 2022) was employed to automatically generate the solution scheme as Julia code, and along with solvers from NonlinearSolve.jl, solve the system of equations in the forward problem. Presently, we have implemented the Julia version in serial mode only. Results presented in this paper were produced using the PyTorch version, although the computational results were the same with the Julia version. Hence, we think both versions have value for future effort.

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f02

Figure 2Model equations corresponding to f and h in Eq. (1). Blue boxes indicate equations corresponding to f. Yellow boxes indicate equations corresponding to h. First, we obtain a solution for Ci (intercellular leaf CO2 pressure) by solving the nonlinear system (f equations) as illustrated in the last (bottom) blue box. Then, we run the h equations forward to compute An (net photosynthesis rate) using Ac, Aj, and Ap as discussed in Sect. 2.2. Details about different variables and parameters included in the f and h equations are provided in Appendix A.

Download

2.3 The parameterization pipeline and model changes

We used multilayer perceptron (MLP) NNs as the parameterization module g in Eq. (3). The purpose of the MLPs is to estimate parameters θ, which are then fed into the photosynthesis module to obtain the net photosynthetic rate (An; Appendix A). The MLPs were trained based on minimization of the loss function – in brief, the goal is to minimize the difference between the solved and observed values of An. As described in Eq. (2), the whole workflow is hereafter referred to as the δpsn model (“delta-photosynthesis model”; the Greek letter δ is selected because the model is programmatically differentiable and δ is often associated with gradients). There may be multiple MLPs used to estimate different parameters in θ, each with different inputs of either continuous or categorical data, and they can all be trained together. Figure 1a shows the general framework for parameter estimation experiments and panel (b) shows the framework of this particular work.

We chose to estimate one or both of two specific parameters in our experiments. The first one is the plant maximum carboxylation rate at 25 C (Vc,max25), which is normally formulated as a PFT-dependent parameter. Although Vc,max25 is hypothesized to be PFT dependent, recent studies have shown that the parameter can vary in space and time and by different species in the PFT as well (Ali et al., 2015; Chen et al., 2022; Qian et al., 2019). Estimating Vc,max25 is not a trivial matter due to its high variability and sensitivity to different factors such as drought, leading some studies to suggest a substitute. For example, Croft et al. (2017) suggested using leaf chlorophyll content as a direct proxy for Vc,max25. Nevertheless, considering this is an initial study applying dPL, we followed the convention and parameterized it based on PFT:

(6) V c , max 25 = NN V ( PFT ) ,

where PFT is the plant functional type category (in one-hot encoding format, which translates each category to a binary vector) and the NN used for parameterization of Vc,max25 is referred to as NNV hereafter.

The second parameterization is for parameter B defined by Clapp and Hornberger (1978), which influences the soil water stress function (βt, where the subscript t indicates transpiration). βt, called “btran” in the CLM code, reflects the impacts of soil wetness on stomatal conductance and ranges from zero (extreme dry conditions causing stomata closure) to one (wet conditions with stomata fully open). In the following, we describe B and βt computations as in CLM4.5 (Oleson et al., 2013). B is purely a function of soil properties and is defined for each soil layer as Bi, where i refers to the soil layer (see Appendix B). Bi equations will later be replaced by our NN-based parameterization scheme as explained in Sect. 2.3.1 because they were originally empirical and may not be optimal at the global scale. B comes into play when calculating the soil water potential ψi using a power-law formulation:

(7) ψ i = ψ sat , i × S i - B i ψ c ,

where ψsat,i is the saturated soil matric potential and Si is the soil wetness, both defined for a specific soil layer (see Appendix B for detailed calculations). The plant wilting factor (wi) is then calculated using ψi and other PFT-dependent parameters (defined in CLM4.5) such as the soil matric potentials for closed stomata ψc and open stomata ψo, which represent the soil water potentials when stomata are fully closed and fully open, respectively, as defined in Eq. (8). The factor wi is also dependent on other factors like the temperature of the soil layer (Ti) relative to the freezing temperature (Tf), the volumetric liquid water content (θliq), the volumetric ice content (θice), and the volumetric water content at saturation (θsat). In our calculations, θice was ignored since both the leaf and the air temperatures in our dataset were above freezing (0 C or 273.15 K) by at least 5 C.

(8) w i = ψ c - ψ i ψ c - ψ o θ sat , i - θ ice , i θ sat , i 1 ; T i > T f - 2 and θ liq , i > 0 0 ; T i T f - 2 or θ liq , i 0 .

Finally, βt can be calculated by aggregating the plant wilting factor (wi) and plant root distribution (ri) across different soil layers based on the PFT as

(9) β t = i w i r i .

2.3.1 Model changes

In the original water limitation function in CLM4.5, the stomata response to soil water potential is based on a linear function between the water potential for stomata openness and closedness (see Eq. 8). In light of the possibility that plants could respond differently to soil water potential depending on plant hydraulic traits (Christoffersen et al., 2016), in this study, we modified the soil water limitation for PFTs so that they could have different shapes. Specifically, we defined PFT-dependent soil water stress, ψi(PFT) ranging from ψc and ψo, depending on the soil water content, which is calculated as follows:

(10) ψ i ( PFT ) = ψ o × S i - B i ( soil , PFT ) ψ c .

Bi is a PFT- and soil-texture-dependent shape parameter (between 0 and 1) estimated as

(11) B i = NN B i ( % sand i , % clay i , F om , i , PFT ) ,

where %sandi, %clayi, and Fom,i respectively represent the percentage of sand, the percentage of clay, and the fraction of organic matter in soil layer i. The PFT-dependent soil water stress, ψi(PFT), is then fed into the plant wilting Eq. (8) as follows:

(12) w i = ψ c - ψ i ( PFT ) ψ c - ψ o = ψ c - max ψ ψ c , ψ o × S i - B i ( soil , PFT ) ψ c - ψ o 1 .

The new shape parameter Bi in Eq. (11) has a different range (between 0 and 1) from the original one defined by Clapp and Hornberger (1978) in Eq. (7), and it varies spatially for different static attributes and for different PFTs as well. The default equations in CLM4.5 for computations of Bi (Appendix B) show that the parameter Bi depends on two attributes, %clayi and Fom,i, which is why they were used in NNBi. To account for the dependence of ψsat,i on %sandi (Appendix B) and its replacement by ψo (see Eqs. 7 and 10), %sandi was also added to NNBi. We also added PFT to NNBi inputs because vegetation may interact with the soil moisture constraint and we want to allow relevant factors to be included, rather than restricting the list of inputs to what was previously used in the literature. Since in NNBi we use quantitative inputs (%sandi, %clayi,Fom,i) along with categorical inputs (PFT), we used the embedding layer in PyTorch, which translates each category to a vector of quantitative variables. This categorical data can then easily be combined with other quantitative inputs we provide to our NN.

Using the original Eq. (7) for computing ψi resulted in a plant wilting factor, wi, equal to one for more than 90 % of the data points across different soil layers. Changing Eq. (7) to the form shown in Eq. (10) helped to express more variability in wi and eventually in the computed soil water stress function (βt).

2.4 Input and observation datasets

2.4.1 Forcing and photosynthesis rates

For meteorological forcings, we used the ERA5 Reanalysis dataset (Muñoz Sabater, 2019), which provides hourly estimates of soil moisture at different soil depths. Soil moisture contributes to the computation of βt (see Appendix B), where the soil wetness S depends on both the soil moisture and the saturated soil moisture.

For observations of photosynthesis, we used data from the leaf gas exchange database (Knauer et al., 2018; Lin et al., 2015), which is a global database of stomatal conductance measurements and leaf-level photosynthetic rates. It incorporates data from several sites around the world in Australia, Europe, USA, and Asia (Fig. 3). We refer to this dataset as Lin15 throughout the rest of this work, with 43 sites chosen whose dates and times of measurements were available. Lin15 covered nine different PFT categories: rainfed crop “Crop R”, Broadleaf evergreen tree tropical “BET Tropical”, Broadleaf evergreen tree temperate “BET Temperate”, C3 grass, C4 grass, Needleleaf evergreen tree boreal “NET Boreal”, Needleleaf evergreen tree temperate “NET Temperate”, Broadleaf deciduous tree temperate “BDT Temperate”, and Broadleaf deciduous shrub temperate “BDS Temperate”. Measurements were taken on the sub-hourly scale but not necessarily at a continuous daily interval. That is why for almost all the sites, data were available on some random days (not necessarily continuous) for one or a few years.

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f03

Figure 3Map of sites available from the leaf gas exchange database (Lin et al., 2015). Different symbols represent different PFTs. The C4 site is highlighted by a thick-bordered hexagon. The marker sizes represent the quantity of data available for each site (map based on matplotlib basemap; Hunter, 2007).

Lin15 also contained forcing variables, including air temperature (T), leaf temperature (Tv), atmospheric pressure (Patm), relative humidity (RH), photosynthetic active radiation (φ) and boundary layer conductance (gb). We used ERA5 to fill in for any missing forcing variables in Lin15. In Eq. (4), Patm and gb were used directly from the dataset, while Ca was estimated using observations of the leaf surface CO2 concentrations, and gs was calculated using the Medlyn conductance model (Medlyn et al., 2011) as explained in Appendix A.

2.4.2 Static attributes

For βt calculations, we used data from Hengl and Wheeler (2018) for the soil organic carbon content at different soil depths, where the conventional Van Bemmelen factor of 1.72 was used to convert to soil organic matter (Fom). Data for sand and clay percentages (%sand, %clay) were obtained from Hengl (2018). Both are global datasets available at 250 m resolution at six different soil depths (0, 10, 30, 60, 100, and 200 cm), which describe five soil layers.

2.4.3 CLM4.5 default parameters

CLM4.5 documentation (Oleson et al., 2013) provides reference values for comparison and equations for both target parameters Vc,max25 and B. For Vc,max25, default values corresponding to each PFT (shown in Table 3) are well documented in CLM4.5 (Chap. 8; Table 8.1). Similarly, for parameters B and βt, their default equations (shown in this work in Appendix B) are provided in the documentation of CLM4.5 as well. We also used other PFT photosynthetic parameters required for βt computations, such as the soil matric potentials for closed stomata, ψc, and open stomata, ψo (see Eqs. 8, 10, 12), and the plant root distribution parameters (see Eq. 9).

2.5 Synthetic data and real data experiments

2.5.1 Case 1: synthetic data

In our synthetic experiments, we assumed values for some parameters to generate synthetic photosynthesis rates that could serve as synthetic training data. Then, we estimated those parameters with NNs while keeping other components unmodified. These experiments were intended to verify the plausibility and efficiency of the dPL framework and to verify the identifiability of parameters.

In the first synthetic case, “Vc,max-only”, the δpsn framework was tested for its ability to accurately retrieve a single PFT-dependent parameter, Vc,max25, using NNV. We used the suggested values for Vc,max25 from CLM4.5 for different PFTs to calculate the synthetic net photosynthetic rates (synthetic training data). For this case the βt values were kept constant (equal to one) for all data points, since we intended to test the retrievability of one parameter.

In the second synthetic case, “Vc,max-B”, we tested simultaneously retrieving both Vc,max25 and B, the latter of which varies spatially for different static attributes. For simplicity, we used only the topsoil layer for this case and excluded the influence of the PFT term; therefore we assumed B1=0.1Fom,1+0.45(%sand1+%clay1) to generate the synthetic data. The plant wilting factor (w1) was then calculated using Eq. (12) and was fed into Eq. (9) to compute the soil water stress function (βt). Since we were using only the topsoil layer, βt was simplified to (βt=w1r1) with a root distribution value for the topsoil layer (r1=1). To retrieve B1, we used NNBi (see Eq. 11) but excluded the PFT term since it was not used in synthesizing B1 values.

For both synthetic runs “Vc,max-only” and “Vc,max-B”, the MLP models were trained concurrently for all PFTs with several data points for each PFT. Moreover, white noise was added to the synthetic values of An with a standard deviation of 5 % of the mean value.

2.5.2 Case 2: real data

Once we confirmed that the model passed the test of the synthetic case (correctly retrieving parameter values which were used to generate the data it was given), it was then applied to a real dataset (Lin et al., 2015) using observation data. This tested whether the model, learning from this dataset for many of the PFTs, could find parameters to better describe photosynthesis data than the literature values. There was no ground truth in this case so we tested multiple formulations to better understand the impacts of allowing more or less flexibility in the estimation and role of each parameter.

We tested several formulations to estimate either one (Vc,max25) or two parameters (Vc,max25 and B) at a time. In essence, we compared allowing either one or two of the parameters to be estimated vs. using the default formulation or values from the original model. For Vc,max25, the default values were those defined in CLM4.5, while for βt, the default equations (Appendix B) were used to obtain its values. Altogether, we trained the following models.

  •  

    Vdef+Bdef: in this case, Vc,max25 took the default values from CLM4.5 and B was calculated using the default equations (Appendix B). This was used as a reference case.

  •  

    Vdef+B: in this formulation, the default Vc,max25 values from CLM4.5 were used while B was estimated using NNBi.

  •  

    V+Bdef: in this formulation, Vc,max25 was estimated using NNV, while B was calculated using the default equations (Appendix B).

  •  

    V+B: in this formulation, we employed both NNV and NNBi. They were trained concurrently to see if this interfered with parameter retrieval.

Representing a real case, Bi was estimated for the ith soil layer based on the static attributes for that layer in the four tested model formulations. Thus, Bi varied both horizontally and vertically for each PFT.

Just as in the synthetic case, the MLPs were shared between all sites. All sites were used to calculate one loss function as in typical ML tasks, with the hope of ensuring the wide applicability of the MLPs and leveraging the synergy between all sites (Fang et al., 2022). In this way, we also hoped to identify parameters that generalize well in space and are applicable at large scales.

The MLPs employed had three layers: an input layer, a single hidden layer, and an output layer. To ensure an output value between 0 and 1 for both Vc,max25 and B parameterizations, sigmoid activation functions were used for both hidden and output layers. Vc,max25 was then rescaled to be within a pre-defined range based on literature values of 20 to 150 µmol m−2 s−1. For the ith soil layer, Bi values were kept between 0 and 1, so with Si ranging between 0.01 and 1 (see Appendix B), the term Si-Bi then had a range of 1 to 100. This ensured that the value of ψi ranged from ψc to ψo (see Eq. 10).

The quantity of available data posed a limitation and did not permit an extensive hyperparameter tuning experiment with a train/validation/test split. Hence, we employed a “lazy” trial and error process with hyperparameters (learning rates and hidden size) using 70 % of the data as training data and 30 % as a validation set just to ensure we had a roughly reasonably performing hyperparameter set. We selected a learning rate of 0.045 and a hidden size equal to the number of inputs (nine for NNV and eight for NNBi). We kept these same hyperparameters when we ran five-fold cross validation with an 80 % : 20 % train : test ratio. In addition, we found that moderately perturbing the hyperparameters resulted in very little change in the performance (see Appendix C). This design was necessary considering the practical limits of the available data, even though this study already represents a large-sample study in the domain of ecosystem modeling.

Two different tests were performed with respect to data splitting: temporal holdout and randomized cross validation – the former test stresses the models' ability to project into the future, while the latter is the typical experiment run in the literature. Due to the irregularity of measurement dates at each location (as mentioned previously in Sect. 2.4.1), the temporal periods for the training and testing datasets varied by location. In the temporal holdout test, for each PFT in each location, the available dates of measurements were recorded. The oldest 80 % of these dates were used for training and the remaining most recent 20 % were used for testing. The temporal holdout test was run for both synthetic and real data experiments. For the randomized cross validation test, as the name implies, the dataset was randomly split into five folds (groups) and each time the model was trained on four folds (80 % of the data points) and tested on the 5th fold (20 % of the data points). This was done a total of five rounds so that all of the data points were used for testing once. The cross validation test was run only for the real data experiments.

We then compared the values of Vc,max25 learned by the V+B model, trained on all data points, against values of Vc,max25 in other data sources (Kattge et al., 2020; Rogers, 2014), which highlights the variability of these parameters. The TRY database (Kattge et al., 2020) has Vc,max25 values defined for several species which can be aggregated to get unique values for each PFT (Table 3). Moreover, we compared our Vc,max25 values to the ones used in different Earth system models (ESMs; Rogers, 2014) for various PFTs, e.g., the Atmosphere-Vegetation Interaction Model (AVIM; Ji, 1995) and the Biosphere Energy Transfer Hydrology scheme (BETHY; Knorr and Heimann, 2001). The comparison enabled us to determine whether the inversely determined values were on the same order of magnitude as previously employed in the literature, and were physically plausible. We expected our values for different PFTs to be at least partially correlated with the ones used in the literature, as they were meant to represent the same physical quantity. A complete disagreement or a different order of magnitude would suggest that our values may not be physically representative. Partial discrepancies would highlight any knowledge gaps. We did not perform a similar comparison between learned and computed Bi values from default equations since the new shape parameter Bi(soil,PFT) (see Eq. 11) is different from the original one and has a different range (between 0 and 1).

2.6 Statistical metrics

To evaluate different experiments and explore the sensitivity of the results to changing different parameters, we chose four different metrics as shown in Table 1. The four metrics were root mean square error (RMSE), bias, Pearson's correlation (COR), and Nash–Sutcliffe efficiency (NSE). Both RMSE and bias measure how far the model simulations are from the observations (and thus the ideal value is 0); however, RMSE is the standard deviation of all errors while bias is calculated as the average. COR measures the linear relationship between both the simulations and the observations, ranging between −1 and 1. NSE measures the relative magnitude of the residual variance relative to the observed data variance (Nash and Sutcliffe, 1970), and has a perfect score of 1. Table 1 shows the formulations of the four metrics and their possible ranges.

Table 1Performance metrics used for evaluation and their possible ranges.

σ refers to the standard deviation, OBS refers to the mean of the observed values, and SIM refers to the mean of the simulated values.

Download Print Version | Download XLSX

3 Results

3.1 Results for synthetic data case

The results of the synthetic experiments showed that our workflow successfully recovered the parameters in both the one-parameter case (“Vc,max-only”, Fig. 4) and the two-parameter case (“Vc,max-B”, Fig. 5). In the one-parameter “Vc,max-only” case, the recovered parameters agreed with the assumed values almost completely for each PFT (Fig. 4a). The model was able to capture the variability in the values of Vc,max25 for different PFTs, where the values ranged from 100.7 µmol m−2 s−1 for rainfed crops (defined as Crop R in CLM4.5) to around 50 µmol m−2 s−1 for C4 grass (Fig. 4a). Moreover, we found nearly complete agreement between the synthetic and recovered net photosynthesis rates (An; Fig. 4b). This single-parameter case demonstrated that the dPL framework and the posited formulation Vc,max25=NNV(PFT) were functional, but (as intended) did not show the effects of parameter interactions.

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f04

Figure 4Single parameter recovery for synthetic data. (a) Comparison of modeled parameter values to literature values by PFT. (b) Actual and modeled net photosynthesis rates for training and testing periods (dashed line indicates the ideal 1 : 1 relationship).

Download

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f05

Figure 5Dual parameter recovery for synthetic data. (a) Comparison of modeled parameter values to literature values by PFT estimated using NNV. (b) Actual and modeled parameter values for B, estimated using NNBi. (c) Actual and modeled parameter values for βt for the topsoil layer. (d) Actual and modeled net photosynthesis rates for training and testing periods. Dashed lines in (b)(d) indicate the ideal 1 : 1 relationship.

Download

With the dual-parameter case, we found a similarly near-complete recovery for Vc,max25 (Fig. 5a) and a near-complete reproduction of simulated photosynthesis (Fig. 5d). However, we noticed a negligible amount of scattering with βt (Fig. 5c), and to a larger extent, with B (Fig. 5b). For all experiments, we verified that the training and testing periods were highly consistent (between green and blue points in the scatterplots). The results indicate that the problem formulation allows for sufficient sensitivity of the net photosynthesis rate with respect to PFT-specific Vc,max25 and the soil water constraint. In addition, Vc,max25 and B influence the photosynthesis rate in different ways so that, along with a large dataset with different combinations of moisture conditions and PFTs, they can be identified simultaneously. This forms the basis of the next stage of the work. The soil moisture parameter identifiability was slightly weakened compared to Vc,max25 because there were more equations involved between B and An, and some of them had parameters in the exponential operators, e.g., ψi=ψoSi-Bi. Mathematically, such a curve can be flat and the gradients can be small in some ranges of S. Mechanistically, An can have reduced sensitivity to B under some conditions. Therefore, we do not expect soil properties to be fully identifiable from photosynthesis data, but the general pattern may still be learnable.

3.2 Results for real data case

3.2.1 Comparisons between candidate formulations

In the test cases employing real datasets, the V+B model (employing both NNV and NNBi) exhibited obvious advantages over the default photosynthesis module in FATES model and the default parameters, as well as the models that learned only one of the parameters (Table 2). For the temporal holdout test, the default CLM4.5 parameters (Vdef+Bdef) led to a lower correlation (0.539), a large bias (1.330 µmol m−2 s−1) and nearly zero NSE (0.001, resulting mainly from the large bias; Table 2a). In particular, the default values appeared to cause an underestimation of the net photosynthetic rate (An) for BET Tropical (Fig. 6a, I) and C3 grass (Fig. 6a, II) but large overestimation for the high-photosynthesis data points of C4 plants (Fig. 6a, III). After allowing B to be learned (Vdef+B), the correlation for testing slightly increased to (0.551), while the bias remained high (0.724 µmol m−2 s−1); it seems that the learning of water stress alone did not address the bias. On the other hand, when we only allowed Vc,max25 to be estimated (V+Bdef), the bias greatly increased (−1.653) and the test NSE was slightly decreased to (0.130). Finally, if we allowed both parameters to be learned (V+B), a decent correlation was obtained (0.757), the bias was the smallest value yet (−0.327µmol m−2 s−1), and the test NSE was 0.565, which means the model explained about half of the variance in the observed photosynthesis rate. The remaining error might be attributable to other untuned parameters, processes related to vegetation states which are not considered by the present model. These issues can be potentially further improved in the future using the differentiable modeling paradigm.

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f06

Figure 6Comparisons of photosynthesis model calibration. Comparing impacts of default and learned parameters by plotting observed vs. simulated An (net photosynthetic rate) values calculated using different candidate models (described by which parameter definitions they use). (a) Impact of learning B with default Vc,max25. (b) Impact of learning Vc,max25 with varying B (either learned alongside V in V+B or defined by the default equations in CLM4.5). The colors represent the results from the four different models, the shapes indicate the PFT groups, and the dotted line in each panel indicates the ideal 1 : 1 relationship. Subscript “def” indicates that the variable was calculated using the default definitions in CLM4.5, while no subscript indicates that the parameter was learned using an NN. Scatterplots were created using the test results from each of the five folds of the cross validation test. For better illustration, only three PFTs are placed in a panel, as indicated by the panel titles. Comparing symbols in the same panel gives insights about the role of estimating B, while comparing left and right panels gives insights about the role of estimating Vc,max25.

Download

Table 2Performance metrics for the candidate models for the Lin15 dataset. In the following, the subscript “def” indicates that the default parameter value from CLM4.5 was used, while parameters without “def” were estimated as an output of a NN (in all cases, V indicates that Vc,max25 was estimated as a function of PFT using NNV and B indicates estimation using NNBi). Part (a) shows the temporal holdout test, where the oldest 80 % of data points were used for training and the most recent 20 % were reserved for testing; part (b) shows the cross validation (five-fold) test.

Download Print Version | Download XLSX

A similar behavior in the performance metrics was observed for the five-fold cross-validation test (see Sect. 2; Table 2b). The cross validation test decreased to a great extent the disparity in the metrics' values between the training and testing datasets (Table 2b). However, contrary to the temporal holdout test, we found a slight improvement in COR (0.596) and NSE (0.137) when B was learned (Vdef+B), while a much higher boost was found in the metrics when Vc,max25 was learned (V+Bdef). This shows the higher impact of learning Vc,max25 on the simulation of An, where the COR and NSE increased to 0.695 and 0.449, respectively, while the bias decreased to −0.478. Similar to the temporal holdout test, the V+B model showed the best metrics in comparison to other models, with the lowest RMSE (4.492) and bias (0.022) values, and the highest COR (0.763) and NSE (0.579) values.

Consistent with the observations of the synthetic experiments, Vc,max25 and B impacted An in different ways. When Vc,max25 was not adjusted, the photosynthesis rates simulated for a number of sites in the high An range (most of them C4 plants) had some substantial overestimation, regardless of whether B had learned or default values (Fig. 6a). It was only after the model also learned Vc,max25 that these high biases were reduced (Fig. 6b). Hence, apparently, the learning reduced the Vc,max25 for these sites compared to the default values. In contrast, learning B mainly corrected the low bias for low simulated An data points (specifically for BET Tropical, C3 grass, and C4 grass; Fig. 6b). A group of sites with underestimations in An has been corrected upward (from yellow to green, bottom points in Fig. 6b) due to a correction in the soil parameter B. Apparently, the original parameters overestimated the water stress for these sites. Learning both parameters together was also effective in reducing overestimations and underestimations in the simulated An for NET Boreal and BDS Temperate, respectively. Our results suggest the adjustments to both parameters improved the results, but Vc,max25 was more impactful, especially in addressing the bias.

We also noticed the different PFTs benefited differently from the parameter learning. For example, BDS Temperate and Crop R did not benefit much (compare red symbols in Fig. 6a and green symbols in Fig. 6b), BET Tropical and NET Temperate saw moderately improved correlation, while C3 grass and C4 grass saw significant improvements in both correlation and bias. These observations indicate the parameters (and thus related processes) tuned here (Vc,max25 and B) have large impacts on C3 and C4 grass while other untuned processes, e.g., vegetation growth and nutrient states, may be contributing to the errors with BDS Temperate and Crop R. C3 plants' improvement is mostly due to learning B, as they are more sensitive to drought in the model, while C4 plants' improvement is due to learning Vc,max25, as they are more resistant to drought but more sensitive to light in the model.

In addition, our test showed that the framework is moderately impacted by long-term nonstationarity, as the temporal test had worse metrics than the cross-validation test (compare Table 2 part b with a). The absolute value of the bias increased from 0.022 in the cross validation test to 0.327 in the temporal test. This suggests that the current model (and perhaps the training data) still has some limitations with representing long-term changes. Possible reasons may include CO2 fertilization and its impact on water use efficiency or differences in the state of plants, as this factor is not included in our present parameterization. In the future, these issues could be addressed by assembling a more long-term training dataset (the Lin15 dataset has data ranging from 1991 to 2013), as well as improving the parameterization and physics of the model.

3.2.2 Recovered parameters

Even though we did not prescribe the values of Vc,max25, the training on the dataset converged to parameter values that were partially correlated with, yet still substantially different from, the literature values, and were on the same order of magnitude (Fig. 7). The default Vc,max25 values came from in situ measurements at a limited number of sites, while our values came from learning from a moderately large dataset (essentially an inversion process limited to the model structure). The fact that they agreed with each other in the main pattern suggests both have merit, and that the learning process captured fundamental physics. The upper half of Fig. 7b saw a high correlation, but Vc,max25 values for the V+B model were uniformly higher than the CLM4.5 defaults, especially for the NET Boreal PFT. The correlation was lower toward the lower half of Fig. 7b (where Vc,max25 from CLM4.5 was lower than 65 µmol m−2 s−1) – the learned Vc,max25 had a larger variability. In particular, the learned Vc,max25 (V+B) for C4 grass is much lower than the default, which could be attributed to species-level variability and the fact that the dataset contains very limited sites with C4 plants. Hence, we do not argue that the values learned here would be applicable globally to other C4 grasses. It seems that the inter-PFT variability in Vc,max25 was previously underrepresented by the CLM4.5 default parameter values (BET Tropical, BET Temperate, BDS Temperate, C4 grass), and the learning process used here enhanced the variability. Moreover, we note that for either Crop R, BET Tropical, BET/BDT Temperate, or C4 grass, the influence of learning B on Vc,max25 was mostly small (Vc,max25 from V+B and V+Bdef models were mostly similar), and thus the interactions between these two parameters were not significant. The overall results showcase the ability of the algorithm to adapt to data and reveal parameter interactions.

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f07

Figure 7Parameter recovery for real data. (a) Comparison of modeled parameter values to literature values of Vc,max25 by PFT. (b) Actual and modeled Vc,max25 values plotted by PFT (dashed line indicates the 1 : 1 ideal relationship). In this figure, both “V+Bdef” and “V+B” models were trained using the whole dataset.

Download

In our interpretation, the learned values represent a more “fine-tuned” version of the literature Vc,max25 values, with the interference from soil water stress disentangled. The magnitude and ranking for PFTs remained similar to the literature values, but the results were improved in different ways for different PFTs. The V+B model obtained lower Vc,max25 for C4 grass, addressing the significant overestimation bias for these sites, which we noted in Fig. 6a, III. Due to their different photosynthesis pathway, C4 plants have the lowest learned Vc,max25, but overall the highest net photosynthesis rates, which were not heavily influenced by the choice of the B parameter. For C3 grass, V+B only slightly increased Vc,max25 compared to the default CLM4.5 values, which addressed the notable low bias in Fig. 6b, II. The default soil parameterization for C3 grass sites seemed somewhat deficient as soil water stress accounted for the other parts of variance in net photosynthesis, as demonstrated by the comparison between V+B and V+Bdef models in Figs. 6b-II and 7b for C3 grass.

We compared our learned Vc,max25 values (Table 3 and Fig. 8) with values from other ESMs and with some observatory values in the TRY database (Kattge et al., 2020; Rogers, 2014). The learned Vc,max25 values are higher than those of the TRY database for most PFT classes except for BDT Temperate and BDS Temperate; however, they are within the range of values used in other ESMs except for relatively higher estimations for Crop R, NET Boreal, and C3 grass. On the scale of ESMs, several values for Vc,max25 are adopted by those models. We computed the correlation coefficient between our learned Vc,max25 values and reference values from other ESMs and the TRY database, finding high correlations (except for AVIM) between the learned and reference Vc,max25 values for CLM4.5 (0.844), BETHY (0.900), and the TRY database (0.699). For instance, Vc,max25 for C4 grass is 25 and 20 (µmol m−2 s−1) in AVIM and BETHY, respectively (Table 3). These values agree with the learned Vc,max25 of the V+B model of 22.90 (µmol m−2 s−1), whereas much higher values were found to be adopted for C4 grass, with 60 (µmol m−2 s−1) being used in the biogeochemical cycles model BiomeBGC as reported in Rogers (2014) and 51.6 (µmol m−2 s−1) in CLM4.5. Vc,max25 from the V+B model and TRY database are similar for BET Tropical and BDT Temperate. For BDS Temperate, the learned Vc,max25 was lower than that in TRY by  20 (µmol m−2 s−1), but similar values were used by BETHY and lower values were used by AVIM. For Crop R, NET Boreal, NET Temperate, and BET Temperate, the learned Vc,max25 values were all  20–30 (µmol m−2 s−1) higher than those of the TRY database, but (except for NET Boreal) similar values have been used in AVIM or BETHY. Both the learned (V+B) and the observed (TRY database) Vc,max25 values show a similar pattern to the lowest Vc,max25 for BET Tropical and a high value assigned for Crop R.

Table 3Vc,max25 simulated by the V+B model versus observed values from the TRY database (with partial overlap in species with the Lin15 dataset – the percentage of overlap is provided in the table), and used in different ESMs such as CLM4.5, AVIM, and BETHY.

Download Print Version | Download XLSX

https://bg.copernicus.org/articles/20/2671/2023/bg-20-2671-2023-f08

Figure 8The correlation between the Vc,max25 values estimated by the V+B model on the y axis versus Vc,max25 values from CLM4.5 (black markers), AVIM (cyan markers), BETHY (magenta markers), and the TRY database (orange markers). Different marker shapes represent different PFTs, while different colors represent different reference sources for Vc,max25 per PFT. For the TRY database, we do not have values for C3 grass and C4 grass due to the lack of overlap in species between the TRY database and our dataset for those two PFTs. The dashed line indicates a 1 : 1 relationship.

Download

4 Discussion

As an initial exploration of the potential of the emerging differentiable computing paradigm (a genre of physics-informed ML; Shen et al., 2023) for applications in ecosystem modeling, our work showed promise but also had many limitations, as the goal was not to produce the best-performing photosynthetic model. We restricted our parameter sets to be dependent on PFT, whereas it is known that within-PFT variation can be significant, and parameters could also be determined on the trait level as well as by multiple environmental factors. Our model did not consider the effects of memory (e.g., rainfall in previous days) or the state of the vegetation (e.g., carbon stored in the canopy or carbon : nitrogen ratios in the canopy). The soil moisture data comes from the ERA5 dataset, which, based on comparisons to in situ data, would be outperformed by ML-based soil moisture predictions (Fang et al., 2017; Liu et al., 2022, 2023), but we used it due to its seamless global coverage and availability for multiple soil depths. This work also only modified the parameterization scheme and did not learn model structures. Recently, development in differentiable hydrologic models has allowed for learning parts of the model using NNs (Feng et al., 2022a, b). In summary, we believe there is still much room for improving model quality, but at some point we will likely run into the limits of measurements (aleatoric uncertainty) or data availability (epistemic uncertainty; Hüllermeier and Waegeman, 2021). Future effort can harness deep networks to establish reference levels as a measure of the data uncertainty (Feng et al., 2022a).

This work appears to be the first evaluation of the Lin15 dataset, and, as such, it establishes a reference level to which future studies can compare. The current dataset may still have limitations in that the number of sites for C4 plants is small and in that it does not allow for ample testing. Some geoscientific domains have well-known benchmark datasets, e.g., the CAMELS dataset in hydrology (Feng et al., 2020). Having such a common (and hopefully large) benchmark dataset allows better model structures to be rapidly discovered and is highly beneficial to the growth of the community (Shen et al., 2018). Related to the limits of measurement errors discussed above, multiple deep-learning-based studies have explored the approximate data error limit (or best achievable model) of CAMELS and that knowledge has been appreciated by the community (Feng et al., 2021). Moreover, deep learning methods benefit from data synergy effects (Fang et al., 2022), where more sites and more diverse data lead to a more robust model and better performance for each site.

Although applying the dPL framework improved the parameters to an extent, the model still has similar structural limitations as other Farquhar-type models. We did not test the model's ability to capture the seasonality of the net photosynthetic rate due the limited site-level temporal data. The seasonal behavior of the model is expected to be similar to other Farquhar models as here we only learned static parameters. Further improvement likely will need to consider vegetation growth. Also, this study did not cover spatial generalization since we do not present results for spatial tests nor is it based on site-level comparison. Improving spatial generalization may require further changes in the model or dynamical parameters or the use of other error mitigation approaches (Feng et al., 2021, 2022b; Ma et al., 2021). This was not within the scope of this study; however, it will be considered for future work.

We would like to highlight that such parameterizations are suitable to the target and forcing datasets used in training (which is still the most representative accessible dataset) and are related to the PBM employed. The dataset may have limitations related to the consistency of the measurement approach, and there may be errors in the forcing data or imperfections in model structure. The model performance may also vary based on the different forcing data and inputs used.

5 Conclusions

In this study, we proposed a novel differentiable ecosystem modeling framework that uses NNs as a parameterization scheme to support a process-based ecosystem model (FATES). Training coupled NNs was not previously possible without differentiable programming, and it allows us to approximate complex a priori unknown mapping relationships between PFTs, landscape characteristics, and physical parameters. The photosynthesis module was treated as a system of nonlinear equations, and, like other such systems, could be solved efficiently and in a massively parallel fashion on GPUs by our differentiable framework. Vc,max25 and a soil water parameter (B) could be simultaneously identified in our synthetic experiments, because they played different roles in the model.

Compared to purely data-driven ML approaches, the differentiable programming framework provides physically meaningful variables and can be used to learn relationships from big data. Via training on a global dataset, we found Vc,max25 values for global sites that correlate with the values in the literature but produce more accurate net photosynthesis rates. It is noteworthy that these values were identified without any supervision from experts other than the preparation of the training dataset and the model. We conclude that Vc,max25 has a larger impact on photosynthesis than the soil water stress parameter, but both can be useful in tuning model responses, with varied impacts on different plants, and their default values were not optimal. Not only is differentiable modeling able to improve simulation quality and provide model parameterization, it also allows us to modify model structure and ask questions regarding unclear parts of the model in the future. There is significant room for this framework to improve and expand to other ecosystem modeling applications.

Appendix A: The system of nonlinear equations

The FATES photosynthesis module is based on the classical Farquhar model for C3 plants (Farquhar et al., 1980), which calculates the photosynthetic rate based on carbon fluxes under different limitations. For C4 plants, it uses the Collatz model (Collatz et al., 1992). Both models assume that the gross photosynthetic rate is affected by the maximum rate of carboxylation and is limited by the concentration of Rubisco (Ac; see Eq. A1), light and electron transport (Aj; see Eq. A2), and the concentration of PEP carboxylase enzyme in C4 plants (Ap; see Eq. A3). Ac,Aj, and Ap are calculated as

(A1)Ac=Vc,max×Ci-ΓCi+Kc1+OiKoforC3plantsVc,maxforC4plants,(A2)Aj=Jx×Ci-Γ4Ci+8ΓforC3plantsα(4.6φ)forC4plants,(A3)AP=KpCiPatmforC4plants,

where Vc,max is the maximum carboxylation rate, Ci is the intercellular leaf CO2  pressure (nonlinear system output), Γ is the CO2 compensation point, Kc and Ko are the Michaelis–Menten constants, Oi is the O2 partial pressure (calculated as 20 % of the atmospheric pressure), Jx is the electron transport rate (see Eqs. A4 and A5), α is the quantum efficiency (0.05 mol CO2 mol−1 photon), φ is the photosynthetically active radiation (available in Lin15), Kp is the initial slope of C4 CO2 response curve, and Patm is the atmospheric pressure (available in Lin15).

Γ, Kc, and Ko are the scaled parameters based on leaf temperature (Tv) calculated using their standardized values at 25 C which are Γ25=42.75×10-6Patm, Kc,25=404.9×10-6Patm, and Ko,25=278.4×10-3Patm, multiplied by the temperature response functions defined in chapter 9.0 in CLM5.0 (Lawrence et al., 2019).

Jx is given by the minimum root of the following quadratic equation:

(A4) θ PSII J x 2 - I PSII + J max J x + I PSII J max = 0 ,

where Jmax is the maximum electron transport rate, θPSII is an empirical curvature for the electron transport rate (0.7) and IPSII is the light utilized in electron transport calculated using a quantum yield parameter (ΦPSII=0.85) as

(A5) I PSII = 0.5 Φ PSII ( 4.6 φ ) .

The three biophysical rates Vc,max, Jmax, and Kp along with the plant respiration rate (Rd), adjusted for Tv, are calculated using their standardized values at 25 C multiplied by temperature response functions defined in Chap. 9.0 in CLM5.0 (Lawrence et al., 2019). Vc,max is also adjusted for the soil water availability by multiplying it with the soil water stress function (βt).

In our case, Vc,max25 is either the default value provided in CLM4.5 or it is learned by an NN, which then is used to calculate other standardized biophysical rates as

(A6)Jmax25=1.67Vc,max25,(A7)Rd25=0.015Vc,max25forC3plants0.025Vc,max25forC4plants,(A8)Kp25=20000Vc,max25forC4plants.

The gross photosynthetic rate (A) is then calculated by solving for the minimum root of the quadratic equations:

(A9)θcjAi2-Ac+AjAi+AcAj=0,(A10)θipA2-Ai+ApA+AiAp=0,

where Ai is an intermediate co-limited photosynthetic rate calculated using the empirical curvature parameter (θcj=0.999). Using Ai, Ap, and the empirical curvature parameter (θip=0.999), the gross rate (A) is given by the smaller root of Eq. (A10). The net photosynthetic rate (An) is

(A11) A n = A - R d .

Then using An, the CO2 partial pressure at the leaf surface (Cs) is calculated as

(A12) C s = C a - 1.4 P atm A n g b 1.0 × 10 - 6 ,

where Ca is CO2 partial pressure near the leaf surface (estimated using observations of the leaf surface CO2 concentrations in Lin15) and gb is the leaf boundary layer conductance, which was available in Lin15 for some sites and the missing values were filled using the mean gb of the whole dataset. The stomatal conductance (gs) is then given by the maximum root of the quadratic equation:

(A13) g s 2 + b g s + c = 0 ,

where b, c, and d are functions in some PFT-dependent parameters:

(A14)b=-2(go+d)+g1d2gb×vpd,(A15)c=go2+2go+d1-g12vpdd,(A16)d=1.6AnCs/Patm.

go, the water stressed minimum stomatal conductance, is calculated as the multiplication of βt and the unstressed minimum stomatal conductance (10 000 µmol m−2 s−1 for C3, 40 000 µmol m−2 s−1 for C4). g1, the slope of the Medlyn stomatal conductance model (Medlyn et al., 2011), is a PFT-specific parameter defined in CLM5.0 (Lawrence et al., 2019). vpd, the vapor pressure deficit, was available in Lin15. Finally, Ci, is related to An using Ca, Patm, gs, and gb as the following:

(A17) C i = C a - A n P atm ( 1.4 g s + 1.6 g b ) ( g s × g b ) .
Appendix B: Computations of btran (βt) in CLM4.5

βt is calculated by aggregating the plant wilting factor (wi) and plant root distribution (ri) across different soil different layers as

(B1) β t = i w i r i .

The plant wilting factor (wi) for soil layer i is mainly dependent on the soil water potential ψi and other PFT-dependent parameters such as the soil matric potentials for closed stomata ψc and open stomata ψo, which represent the soil water potentials when stomata are fully closed and fully open, respectively. The factor wi is also dependent on other factors like the temperature of the soil layer (Ti) relative to the freezing temperature (Tf), the volumetric liquid water content (θliq,i) and volumetric ice (θice,i) content, and the volumetric water content at saturation (θsat,i).

(B2) w i = ψ c - ψ i ψ c - ψ o θ sat , i - θ ice , i θ sat , i 1 ; T i > T f - 2 and θ liq , i > 0 0 ; T i T f - 2 or θ liq , i 0 .

The soil matric potential ψi is calculated using a power-law formulation:

(B3) ψ i = ψ sat , i × S i - B i ψ c ,

where ψsat,i is the saturated soil matric potential, Si is the soil wetness, and Bi is the Clapp–Hornberger parameter, all defined for a specific soil layer (i). Different soil attributes such as percentages of sand (%sandi) and clay (%clayi), fraction of organic matter (Fom,i), and soil moisture (θliq,i) are used in computing ψsat,i, Si, and Bi. ψsat,i is calculated as

(B4) ψ sat , i = 1 - F om , i × ψ sat , min , i + F om , i × ψ sat , om ,

where ψsat,om is the saturated organic matter matric potential (−10.3 mm; Letts et al., 2000) and ψsat,min,i is the saturated mineral soil matric potential calculated using %sandi as

(B5) ψ sat , min , i = - 10.0 × 10 1.88 - 0.0131 × ( % sand ) i .

The soil wetness (Si) is calculated using the volumetric contents θliq,i, θice,i, and θsat,i as

(B6) S i = θ liq , i θ sat , i - θ ice , i , 0.01 S i 1 ,

where θsat,i for a soil layer is

(B7) θ sat , i = 1 - F om , i × θ sat , min , i + F om , i × θ sat , om .

θsat,om is the porosity of the organic matter (0.9; Letts et al., 2000; Farouki, 1981), while the porosity of the mineral soil (θsat,min) using %sand is

(B8) θ sat , min , i = 0.489 - 0.00126 × ( % sand ) i .

Similar to ψsat,i and θsat,i (see Eqs. B4 and B7), the parameter Bi is calculated as

(B9) B i = 1 - F om , i × B min , i + F om , i × B om ,

where Bom is the parameter for organic matter (2.7; Letts et al., 2000), while Bmin,i, the parameter for mineral soil, is

(B10) B min , i = 2.91 + 0.159 × ( % clay ) i .
Appendix C

Table C1V+B model formulation performance for different sizes of NNBi [input size, hidden size, output size] with a 80 % : 20 % train : test split ratio.

* In the last experiment we used two hidden layers (i.e., NNBi [input size, hidden layer1 size, hidden layer2 size, output size]).

Download Print Version | Download XLSX

Code availability

A code example demonstrating the differentiable model and its training process is available at https://github.com/hydroPKDN/diffEcosys/ (last access: 21 June 2023) and citable via https://doi.org/10.5281/zenodo.8067204 (Aboelyazeed et al., 2023).

Data availability

The datasets used in the model are publicly available from the sources cited in this paper. The leaf gas exchange database (Knauer et al., 2018; Lin et al., 2015) can be accessed at (https://bitbucket.org/gsglobal/leafgasexchange, last access: 15 May 2023)

Author contributions

DA implemented the numerical models, ran experiments, and produced the figures. DA and CS completed the initial manuscript. CS and CX conceived the study. CX, FMH, KL, and CS edited the manuscript. JL contributed to data downloading and processing. CS implemented the parallel Newton solver for the nonlinear system in PyTorch, while DA implemented the Julia version with assistance from AWJ and CR.

Competing interests

Kathryn Lawson and Chaopeng Shen have financial interests in HydroSapient, Inc., a company which could potentially benefit from the results of this research. This interest has been reviewed by The Pennsylvania State University in accordance with its individual conflict of interest policy, for the purpose of maintaining the objectivity and the integrity of research at The Pennsylvania State University.

Disclaimer

Publisher’s note: Copernicus Publications remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Acknowledgements

We thank the providers of the datasets for their dedication in ensuring the accessibility of the data to the public. Additionally, we express our gratitude to the reviewers and editor for their valuable feedback, which has greatly contributed to improving the quality of our work.

Financial support

This work was supported by the U.S. Department of Energy, Office of Science, under award no. DE-SC0021979. Chonggang Xu was supported by the Department of Energy, Office of Science, project Next Generation Ecosystem Experiment-Tropics (NGEE-Tropics (award no. DE-SC7537545)).

Review statement

This paper was edited by Andreas Ibrom and reviewed by two anonymous referees.

References

Aboelyazeed, D., Xu, C., Hoffman, F. M., Liu, J., Jones, A. W., Rackauckas, C., Lawson, K. E., and Shen, C.: A differentiable, physics-informed ecosystem modeling and learning framework for large-scale inverse problems, Zenodo [code], https://doi.org/10.5281/zenodo.8067204, 2023. 

Ali, A. A., Xu, C., Rogers, A., McDowell, N. G., Medlyn, B. E., Fisher, R. A., Wullschleger, S. D., Reich, P. B., Vrugt, J. A., Bauerle, W. L., Santiago, L. S., and Wilson, C. J.: Global-scale environmental control of plant photosynthetic capacity., Ecol. Appl., 25, 2349–2365, https://doi.org/10.1890/14-2111.1, 2015. 

Ali, A. A., Xu, C., Rogers, A., Fisher, R. A., Wullschleger, S. D., Massoud, E. C., Vrugt, J. A., Muss, J. D., McDowell, N. G., Fisher, J. B., Reich, P. B., and Wilson, C. J.: A global scale mechanistic model of photosynthetic capacity (LUNA V1.0), Geosci. Model Dev., 9, 587–606, https://doi.org/10.5194/gmd-9-587-2016, 2016. 

Baydin, A. G., Pearlmutter, B. A., Radul, A. A., and Siskind, J. M.: Automatic differentiation in machine learning: A survey, J. Mach. Learn. Res., 18, 1–43, 2018. 

Beven, K.: A manifesto for the equifinality thesis, J. Hydrol., 320, 18–36, https://doi.org/10/ccx2ks, 2006. 

Beven, K. and Freer, J.: Equifinality, data assimilation, and uncertainty estimation in mechanistic modelling of complex environmental systems using the GLUE methodology, J. Hydrol., 249, 11–29, https://doi.org/10/fgmngv, 2001. 

Bezanson, J., Karpinski, S., Shah, V. B., and Edelman, A.: Julia: A fast dynamic language for technical computing, arXiv [preprint], https://doi.org/10.48550/arXiv.1209.5145, 24 September 2012. 

Chen, J. M., Wang, R., Liu, Y., He, L., Croft, H., Luo, X., Wang, H., Smith, N. G., Keenan, T. F., Prentice, I. C., Zhang, Y., Ju, W., and Dong, N.: Global datasets of leaf photosynthetic capacity for ecological and earth system research, Earth Syst. Sci. Data, 14, 4077–4093, https://doi.org/10.5194/essd-14-4077-2022, 2022. 

Christoffersen, B. O., Gloor, M., Fauset, S., Fyllas, N. M., Galbraith, D. R., Baker, T. R., Kruijt, B., Rowland, L., Fisher, R. A., Binks, O. J., Sevanto, S., Xu, C., Jansen, S., Choat, B., Mencuccini, M., McDowell, N. G., and Meir, P.: Linking hydraulic traits to tropical forest function in a size-structured and trait-driven model (TFS v.1-Hydro), Geosci. Model Dev., 9, 4227–4255, https://doi.org/10.5194/gmd-9-4227-2016, 2016. 

Clapp, R. B. and Hornberger, G. M.: Empirical equations for some soil hydraulic properties, Water Resour. Res., 14, 601–604, https://doi.org/10.1029/wr014i004p00601, 1978. 

Collatz, G., Ribas-Carbo, M., and Berry, J.: Coupled Photosynthesis-Stomatal Conductance Model for Leaves of C4 Plants, Aust. J. Plant Physiol., 19, 519, https://doi.org/10/cw8rtn, 1992. 

Croft, H., Chen, J. M., Luo, X., Bartlett, P., Chen, B., and Staebler, R. M.: Leaf chlorophyll content as a proxy for leaf photosynthetic capacity, Glob. Change Biol., 23, 3513–3524, https://doi.org/10.1111/gcb.13599, 2017. 

Dusenge, M. E., Duarte, A. G., and Way, D. A.: Plant carbon metabolism and climate change: elevated CO2 and temperature impacts on photosynthesis, photorespiration and respiration, New Phytol., 221, 32–49, https://doi.org/10.1111/nph.15283, 2019. 

ElSaadani, M., Habib, E., Abdelhameed, A. M., and Bayoumi, M.: Assessment of a Spatiotemporal Deep Learning Approach for Soil Moisture Prediction and Filling the Gaps in Between Soil Moisture Observations, Fr. Art. Int., 4, 636234, https://doi.org/10.3389/frai.2021.636234, 2021. 

Fang, K. and Shen, C.: Near-real-time forecast of satellite-based soil moisture using long short-term memory with an adaptive data integration kernel, J. Hydrometeor., 21, 399–413, https://doi.org/10.1175/jhm-d-19-0169.1, 2020. 

Fang, K., Shen, C., Kifer, D., and Yang, X.: Prolongation of SMAP to spatiotemporally seamless coverage of continental U.S. using a deep learning neural network, Geophys. Res. Lett., 44, 11030–11039, https://doi.org/10.1002/2017gl075619, 2017. 

Fang, K., Kifer, D., Lawson, K., Feng, D., and Shen, C.: The data synergy effects of time-series deep learning models in hydrology, Water Resour. Res., 58, e2021WR029583, https://doi.org/10.1029/2021WR029583, 2022. 

Farouki, O. T.: The thermal properties of soils in cold regions, Cold Reg. Sci. Technol., 5, 67–75, https://doi.org/10.1016/0165-232X(81)90041-0, 1981. 

Farquhar, G. D. and von Caemmerer, S.: Modelling of Photosynthetic Response to Environmental Conditions, in: Physiological Plant Ecology II: Water Relations and Carbon Assimilation, edited by: Lange, O. L., Nobel, P. S., Osmond, C. B., and Ziegler, H., Springer Berlin Heidelberg, Berlin, Heidelberg, 549–587, https://doi.org/10.1007/978-3-642-68150-9_17, 1982. 

Farquhar, G. D., von Caemmerer, S., and Berry, J. A.: A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species, Planta, 149, 78–90, https://doi.org/10/fs9dpz, 1980. 

Feng, D., Fang, K., and Shen, C.: Enhancing streamflow forecast and extracting insights using long-short term memory networks with data integration at continental scales, Water Resour. Res., 56, e2019WR026793, https://doi.org/10.1029/2019WR026793, 2020. 

Feng, D., Lawson, K., and Shen, C.: Mitigating prediction error of deep learning streamflow models in large data-sparse regions with ensemble modeling and soft data, Geophys. Res. Lett., 48, e2021GL092999, https://doi.org/10.1029/2021GL092999, 2021. 

Feng, D., Liu, J., Lawson, K., and Shen, C.: Differentiable, learnable, regionalized process-based models with multiphysical outputs can approach state-of-the-art hydrologic prediction accuracy, Water Resour. Res., 58, e2022WR032404, https://doi.org/10.1029/2022WR032404, 2022a. 

Feng, D., Beck, H., Lawson, K., and Shen, C.: The suitability of differentiable, learnable hydrologic models for ungauged regions and climate change impact assessment, Hydrol. Earth Syst. Sci. Discuss. [preprint], https://doi.org/10.5194/hess-2022-245, in review, 2022b. 

Fisher, R. A., Muszala, S., Verteinstein, M., Lawrence, P., Xu, C., McDowell, N. G., Knox, R. G., Koven, C., Holm, J., Rogers, B. M., Spessa, A., Lawrence, D., and Bonan, G.: Taking off the training wheels: the properties of a dynamic vegetation model without climate envelopes, CLM4.5(ED), Geosci. Model Dev., 8, 3593–3619, https://doi.org/10.5194/gmd-8-3593-2015, 2015. 

Fisher, R. A., Koven, C. D., Anderegg, W. R. L., Christoffersen, B. O., Dietze, M. C., Farrior, C. E., Holm, J. A., Hurtt, G. C., Knox, R. G., Lawrence, P. J., Lichstein, J. W., Longo, M., Matheny, A. M., Medvigy, D., Muller-Landau, H. C., Powell, T. L., Serbin, S. P., Sato, H., Shuman, J. K., Smith, B., Trugman, A. T., Viskari, T., Verbeeck, H., Weng, E., Xu, C., Xu, X., Zhang, T., and Moorcroft, P. R.: Vegetation demographics in Earth System Models: A review of progress and priorities, Glob. Change Biol., 24, 35–54, https://doi.org/10.1111/gcb.13910, 2018. 

Gowda, S., Ma, Y., Cheli, A., Gwóÿzdÿ, M., Shah, V. B., Edelman, A., and Rackauckas, C.: High-performance symbolic-numerics via multiple dispatch, ACM Commun. Comput. Algebra, 55, 92–96, https://doi.org/10.1145/3511528.3511535, 2022. 

Hengl, T.: Sand content in % (kg/kg) at 6 standard depths (0, 10, 30, 60, 100 and 200 cm) at 250 m resolution (v0.2), Zenodo [data set], https://doi.org/10.5281/ZENODO.2525662, 2018. 

Hengl, T. and Wheeler, I.: Soil Organic Carbon Content In X 5 G/Kg At 6 Standard Depths (0, 10, 30, 60, 100 And 200 Cm) At 250 M Resolution, Zenodo [data set], https://doi.org/10.5281/ZENODO.1475458, 2018. 

Hossain, M. S., Al-Hammadi, M., and Muhammad, G.: Automatic fruit classification using deep learning for industrial applications, IEEE T. Ind. Inform., 15, 1027–1034, https://doi.org/10.1109/TII.2018.2875149, 2019. 

Hrnjica, B., Mehr, A. D., Jakupoviæ, E., Crnkiæ, A., and Hasanagiæ, R.: Application of deep learning neural networks for nitrate prediction in the Klokot River, Bosnia and Herzegovina, in: 2021 7th International Conference on Control, Instrumentation and Automation (ICCIA), 2021 7th International Conference on Control, Instrumentation and Automation (ICCIA), 1–6, https://doi.org/10.1109/ICCIA52082.2021.9403565, 2021. 

Hüllermeier, E. and Waegeman, W.: Aleatoric and epistemic uncertainty in machine learning: an introduction to concepts and methods, Mach. Learn., 110, 457–506, https://doi.org/10.1007/s10994-021-05946-3, 2021. 

Hunter, J. D.: Matplotlib: A 2D graphics environment, Comput. Sci. Engin., 9, 90–95, https://doi.org/10.1109/MCSE.2007.55, 2007. 

Innes, M., Edelman, A., Fischer, K., Rackauckas, C., Saba, E., Shah, V. B., and Tebbutt, W.: A Differentiable Programming System to Bridge Machine Learning and Scientific Computing, arXiv [preprint], https://doi.org/10.48550/arXiv.1907.07587, 18 July 2019. 

Ji, J.: A climate-vegetation interaction model: Simulating physical and biological processes at the surface, J. Biogeogr., 22, 445–451, https://doi.org/10.2307/2845941, 1995. 

Kattge, J., Bönisch, G., Díaz, S., Lavorel, S., Prentice, I. C., Leadley, P., Tautenhahn, S., Werner, G. D. A., Aakala, T., Abedi, M., Acosta, A. T. R., Adamidis, G. C., Adamson, K., Aiba, M., Albert, C. H., Alcántara, J. M., Alcázar C, C., Aleixo, I., Ali, H., Amiaud, B., Ammer, C., Amoroso, M. M., Anand, M., Anderson, C., Anten, N., Antos, J., Apgaua, D. M. G., Ashman, T.-L., Asmara, D. H., Asner, G. P., Aspinwall, M., Atkin, O., Aubin, I., Baastrup-Spohr, L., Bahalkeh, K., Bahn, M., Baker, T., Baker, W. J., Bakker, J. P., Baldocchi, D., Baltzer, J., Banerjee, A., Baranger, A., Barlow, J., Barneche, D. R., Baruch, Z., Bastianelli, D., Battles, J., Bauerle, W., Bauters, M., Bazzato, E., Beckmann, M., Beeckman, H., Beierkuhnlein, C., Bekker, R., Belfry, G., Belluau, M., Beloiu, M., Benavides, R., Benomar, L., Berdugo-Lattke, M. L., Berenguer, E., Bergamin, R., Bergmann, J., Bergmann Carlucci, M., Berner, L., Bernhardt-Römermann, M., Bigler, C., Bjorkman, A. D., Blackman, C., Blanco, C., Blonder, B., Blumenthal, D., Bocanegra-González, K. T., Boeckx, P., Bohlman, S., Böhning-Gaese, K., Boisvert-Marsh, L., Bond, W., Bond-Lamberty, B., Boom, A., Boonman, C. C. F., Bordin, K., Boughton, E. H., Boukili, V., Bowman, D. M. J. S., Bravo, S., Brendel, M. R., Broadley, M. R., Brown, K. A., Bruelheide, H., Brumnich, F., Bruun, H. H., Bruy, D., Buchanan, S. W., Bucher, S. F., Buchmann, N., Buitenwerf, R., Bunker, D. E., et al.: TRY plant trait database – enhanced coverage and open access, Glob. Change Biol., 26, 119–188, https://doi.org/10.1111/gcb.14904, 2020. 

Kirschbaum, M. U. F.: Direct and indirect climate change effects on photosynthesis and transpiration, Plant Biol., 6, 242–253, https://doi.org/10.1055/s-2004-820883, 2004. 

Knauer, J., Zaehle, S., Medlyn, B. E., Reichstein, M., Williams, C. A., Migliavacca, M., De Kauwe, M. G., Werner, C., Keitel, C., Kolari, P., Limousin, J.-M., and Linderson, M.-L.: Towards physiologically meaningful water-use efficiency estimates from eddy covariance data, Glob. Change Biol., 24, 694–710, https://doi.org/10.1111/gcb.13893, 2018. 

Knorr, W. and Heimann, M.: Uncertainties in global terrestrial biosphere modeling, Part II: Global constraints for a process-based vegetation model, Global Biogeochem. Cy., 15, 227–246, https://doi.org/10.1029/1998GB001060, 2001. 

Koven, C. D., Knox, R. G., Fisher, R. A., Chambers, J. Q., Christoffersen, B. O., Davies, S. J., Detto, M., Dietze, M. C., Faybishenko, B., Holm, J., Huang, M., Kovenock, M., Kueppers, L. M., Lemieux, G., Massoud, E., McDowell, N. G., Muller-Landau, H. C., Needham, J. F., Norby, R. J., Powell, T., Rogers, A., Serbin, S. P., Shuman, J. K., Swann, A. L. S., Varadharajan, C., Walker, A. P., Wright, S. J., and Xu, C.: Benchmarking and parameter sensitivity of physiological and vegetation dynamics using the Functionally Assembled Terrestrial Ecosystem Simulator (FATES) at Barro Colorado Island, Panama, Biogeosciences, 17, 3017–3044, https://doi.org/10.5194/bg-17-3017-2020, 2020. 

Lawrence, D. M., Fisher, R. A., Koven, C. D., Oleson, K. W., Swenson, S. C., Bonan, G., Collier, N., Ghimire, B., van Kampenhout, L., Kennedy, D., Kluzek, E., Lawrence, P. J., Li, F., Li, H., Lombardozzi, D., Riley, W. J., Sacks, W. J., Shi, M., Vertenstein, M., Wieder, W. R., Xu, C., Ali, A. A., Badger, A. M., Bisht, G., van den Broeke, M., Brunke, M. A., Burns, S. P., Buzan, J., Clark, M., Craig, A., Dahlin, K., Drewniak, B., Fisher, J. B., Flanner, M., Fox, A. M., Gentine, P., Hoffman, F., Keppel-Aleks, G., Knox, R., Kumar, S., Lenaerts, J., Leung, L. R., Lipscomb, W. H., Lu, Y., Pandey, A., Pelletier, J. D., Perket, J., Randerson, J. T., Ricciuto, D. M., Sanderson, B. M., Slater, A., Subin, Z. M., Tang, J., Thomas, R. Q., Val Martin, M., and Zeng, X.: The Community Land Model Version 5: Description of New Features, Benchmarking, and Impact of Forcing Uncertainty, J. Adv. Model. Earth Sy., 11, 4245–4287, https://doi.org/10.1029/2018ms001583, 2019. 

LeCun, Y., Bengio, Y., and Hinton, G.: Deep Learning, Nature, 521, 436–444, https://doi.org/10.1038/nature14539, 2015. 

Leong, W. J. and Horgan, H. J.: DeepBedMap: a deep neural network for resolving the bed topography of Antarctica, The Cryosphere, 14, 3687–3705, https://doi.org/10.5194/tc-14-3687-2020, 2020. 

Letts, M. G., Roulet, N. T., Comer, N. T., Skarupa, M. R., and Verseghy, D. L.: Parametrization of peatland hydraulic properties for the Canadian land surface scheme, Atmos. Ocean, 38, 141–160, https://doi.org/10.1080/07055900.2000.9649643, 2000. 

Lin, Y.-S., Medlyn, B. E., Duursma, R. A., Prentice, I. C., Wang, H., Baig, S., Eamus, D., de Dios, V. R., Mitchell, P., Ellsworth, D. S., de Beeck, M. O., Wallin, G., Uddling, J., Tarvainen, L., Linderson, M.-L., Cernusak, L. A., Nippert, J. B., Ocheltree, T. W., Tissue, D. T., Martin-StPaul, N. K., Rogers, A., Warren, J. M., De Angelis, P., Hikosaka, K., Han, Q., Onoda, Y., Gimeno, T. E., Barton, C. V. M., Bennie, J., Bonal, D., Bosc, A., Löw, M., Macinins-Ng, C., Rey, A., Rowland, L., Setterfield, S. A., Tausz-Posch, S., Zaragoza-Castells, J., Broadmeadow, M. S. J., Drake, J. E., Freeman, M., Ghannoum, O., Hutley, L. B., Kelly, J. W., Kikuzawa, K., Kolari, P., Koyama, K., Limousin, J.-M., Meir, P., Lola da Costa, A. C., Mikkelsen, T. N., Salinas, N., Sun, W., and Wingate, L.: Optimal stomatal behaviour around the world, Nat. Clim. Change, 5, 459–464, https://doi.org/10.1038/nclimate2550, 2015. 

Lin, Y.-S., Medlyn, B., Duursma, R., and Knauer, J.: Leaf Gas Exchange Database, Bitbucket [data set], https://bitbucket.org/gsglobal/leafgasexchange/src/master/, 2017. 

Liu, J., Rahmani, F., Lawson, K., and Shen, C.: A multiscale deep learning model for soil moisture integrating satellite and in situ data, Geophys. Res. Lett., 49, e2021GL096847, https://doi.org/10.1029/2021GL096847, 2022. 

Liu, J., Hughes, D., Rahmani, F., Lawson, K., and Shen, C.: Evaluating a global soil moisture dataset from a multitask model (GSM3 v1.0) with potential applications for crop threats, Geoscientific Model Development, 16, 1553–1567, https://doi.org/10.5194/gmd-16-1553-2023, 2023. 

Luo, X., Keenan, T., Chen, J., Croft, H., Prentice, I., Smith, N., Walker, A., Wang, H., Wang, R., Xu, C., and Zhang, Y.: Global variation in the fraction of leaf nitrogen allocated to photosynthesis, Nat. Commun., 12, 4866, https://doi.org/10.1038/s41467-021-25163-9, 2021. 

Ma, K., Feng, D., Lawson, K., Tsai, W.-P., Liang, C., Huang, X., Sharma, A., and Shen, C.: Transferring hydrologic data across continents – Leveraging data-rich regions to improve hydrologic prediction in data-sparse regions, Water Resour. Res., 57, e2020WR028600, https://doi.org/10.1029/2020wr028600, 2021. 

Ma, Y., Gowda, S., Anantharaman, R., Laughman, C., Shah, V. B., and Rackauckas, C.: ModelingToolkit: A composable graph transformation system for equation-based modeling, CoRR, arXiv [preprint], https://doi.org/10.48550/arXiv.2103.05244, 9 February 2022. 

Mäkelä, J., Knauer, J., Aurela, M., Black, A., Heimann, M., Kobayashi, H., Lohila, A., Mammarella, I., Margolis, H., Markkanen, T., Susiluoto, J., Thum, T., Viskari, T., Zaehle, S., and Aalto, T.: Parameter calibration and stomatal conductance formulation comparison for boreal forests with adaptive population importance sampler in the land surface model JSBACH, Geosci. Model Dev., 12, 4075–4098, https://doi.org/10.5194/gmd-12-4075-2019, 2019. 

Medlyn, B. E., Duursma, R. A., Eamus, D., Ellsworth, D. S., Prentice, I. C., Barton, C. V. M., Crous, K. Y., de Angelis, P., Freeman, M., and Wingate, L.: Reconciling the optimal and empirical approaches to modelling stomatal conductance, Glob. Change Biol., 17, 2134–2144, https://doi.org/10.1111/j.1365-2486.2010.02375.x, 2011. 

Meyer, F. H.: Encyclopedia of Plant Physiology, New Series. Editors: Pirson, A.; Zimmermann, M.H., Vol. 12, Part A (in 4 parts) Physiological Plant Ecology I. Responses to the Physical Environment, Editors: Lange, O.L.; Nobel, P.S.; Osmond, C.B.; Ziegler, H., Springer-Verlag, Berlin–Heidelberg–New York, 1981, 110 figs. XV, 625 pages. Cloth DM 239,–, Z. Pflanz. Bodenkunde, 146, 543–544, https://doi.org/10.1002/jpln.19831460417, 1983. 

Muñoz Sabater, J.: ERA5-Land hourly data from 1950 to present, Copernicus Climate Change Service (C3S) Climate Data Store (CDS) [data set], https://doi.org/10.24381/cds.e2161bac, 2019. 

Nash, J. E. and Sutcliffe, J. V.: River flow forecasting through conceptual models part I – A discussion of principles, J. Hydrol., 10, 282–290, https://doi.org/10/fbg9tm, 1970. 

Oleson, K., Lawrence, D., Bonan, G., Drewniak, B., Huang, M., Koven, C., Levis, S., Li, F., Riley, W., Subin, Z., Swenson, S., Thornton, P., Bozbiyik, A., Fisher, R., Heald, C., Kluzek, E., Lamarque, J.-F., Lawrence, P., Leung, L., Lipscomb, W., Muszala, S., Ricciuto, D., Sacks, W., Sun, Y., Tang, J., and Yang, Z.-L.: Technical description of version 4.5 of the Community Land Model (CLM), UCAR/NCAR, https://doi.org/10.5065/D6RR1W7M, 2013. 

Paszke, A., Gross, S., Massa, F., Lerer, A., Bradbury, J., Chanan, G., Killeen, T., Lin, Z., Gimelshein, N., Antiga, L., Desmaison, A., Kopf, A., Yang, E., DeVito, Z., Raison, M., Tejani, A., Chilamkurthy, S., Steiner, B., Fang, L., Bai, J., and Chintala, S.: PyTorch: An imperative style, high-performance deep learning library, Adv. Neur. In., 32, 8024–8035, 2019. 

Qian, X., Liu, L., Croft, H., and Chen, J.: C3 plants converge on a universal relationship between leaf maximum carboxylation rate and chlorophyll content, Biogeosciences Discuss. [preprint], https://doi.org/10.5194/bg-2019-228, 2019. 

Quillet, A., Peng, C., and Garneau, M.: Toward dynamic global vegetation models for simulating vegetation–climate interactions and feedbacks: recent developments, limitations, and future challenges, Environ. Rev., 18, 333–353, https://doi.org/10.1139/A10-016, 2010. 

Rahmani, F., Oliver, S., Ouyang, W., Appling, A., Lawson, K., and Shen, C.: Developing and testing a long short-term memory stream temperature model in daily and continental scale, AGU 2020 Fall Meeting, Earth and Space Science Open Archive, https://doi.org/10.1002/essoar.10505077.1, 2020. 

Rahmani, F., Shen, C., Oliver, S., Lawson, K., and Appling, A.: Deep learning approaches for improving prediction of daily stream temperature in data-scarce, unmonitored, and dammed basins, Hydrol. Process., 35, e14400, https://doi.org/10.1002/hyp.14400, 2021. 

Reichstein, M., Camps-Valls, G., Stevens, B., Jung, M., Denzler, J., Carvalhais, N., and Prabhat: Deep learning and process understanding for data-driven Earth system science, Nature, 566, 195–204, https://doi.org/10.1038/s41586-019-0912-1, 2019. 

Rogers, A.: The use and misuse of V(c,max) in Earth System Models, Photosynth. Res., 119, 15–29, https://doi.org/10.1007/s11120-013-9818-1, 2014. 

Saha, G. K., Rahmani, F., Shen, C., Li, L., and Cibin, R.: A deep learning-based novel approach to generate continuous daily stream nitrate concentration for nitrate data-sparse watersheds, Sci. Total Environ., 878, 162930, https://doi.org/10.1016/j.scitotenv.2023.162930, 2023. 

Saleem, M. H., Potgieter, J., and Arif, K. M.: Plant disease detection and classification by deep learning, Plants, 8, 468, https://doi.org/10.3390/plants8110468, 2019. 

Shen, C.: Deep Learning: A Next-Generation Big-Data Approach for Hydrology, Eos, 99, https://doi.org/10.1029/2018EO095649, 2018. 

Shen, C., Laloy, E., Elshorbagy, A., Albert, A., Bales, J., Chang, F.-J., Ganguly, S., Hsu, K.-L., Kifer, D., Fang, Z., Fang, K., Li, D., Li, X., and Tsai, W.-P.: HESS Opinions: Incubating deep-learning-powered hydrologic science advances as a community, Hydrol. Earth Syst. Sci., 22, 5639–5656, https://doi.org/10.5194/hess-22-5639-2018, 2018. 

Shen, C., Appling, A. P., Gentine, P., Bandai, T., Gupta, H., Tartakovsky, A., Baity-Jesi, M., Fenicia, F., Kifer, D., Li, L., Liu, X., Ren, W., Zheng, Y., Harman, C. J., Clark, M., Farthing, M., Feng, D., Kumar, P., Aboelyazeed, D., Rahmani, F., Beck, H. E., Bindas, T., Dwivedi, D., Fang, K., Höge, M., Rackauckas, C., Roy, T., Xu, C., and Lawson, K.: Differentiable modeling to unify machine learning and physical models and advance Geosciences, arXiv [preprint], https://doi.org/10.48550/arXiv.2301.04027, 10 January 2023. 

Tang, J. and Zhuang, Q.: Equifinality in parameterization of process-based biogeochemistry models: A significant uncertainty source to the estimation of regional carbon dynamics, J. Geophys. Res., 113, G04010, https://doi.org/10.1029/2008JG000757, 2008. 

Thompson, M., Gamage, D., Hirotsu, N., Martin, A., and Seneweera, S.: Effects of elevated carbon dioxide on photosynthesis and carbon partitioning: A perspective on root sugar sensing and hormonal crosstalk, Front. Physiol., 8, 578, https://doi.org/10.3389/fphys.2017.00578, 2017. 

Tsai, W.-P., Feng, D., Pan, M., Beck, H., Lawson, K., Yang, Y., Liu, J., and Shen, C.: From calibration to parameter learning: Harnessing the scaling effects of big data in geoscientific modeling, Nat. Commun., 12, 5988, https://doi.org/10.1038/s41467-021-26107-z, 2021. 

Urban, L., Aarrouf, J., and Bidel, L.: Assessing the effects of water deficit on photosynthesis using parameters derived from measurements of leaf gas exchange and of chlorophyll A fluorescence, Front. Plant Sci., 8, 2068, https://doi.org/10.3389/fpls.2017.02068, 2017. 

Verheijen, L. M., Brovkin, V., Aerts, R., Bönisch, G., Cornelissen, J. H. C., Kattge, J., Reich, P. B., Wright, I. J., and van Bodegom, P. M.: Impacts of trait variation through observed trait–climate relationships on performance of an Earth system model: a conceptual analysis, Biogeosciences, 10, 5497–5515, https://doi.org/10.5194/bg-10-5497-2013, 2013. 

Von Caemmerer, S.: C4 photosynthesis in a single C3 cell is theoretically inefficient but may ameliorate internal CO2 diffusion limitations of C3 leaves, Plant Cell Environ., 26, 1191–1197, https://doi.org/10.1046/j.0016-8025.2003.01061.x, 2003.  

Von Caemmerer, S.: Steady-state models of photosynthesis, Plant Cell Environ., 36, 1617–1630, https://doi.org/10.1111/pce.12098, 2013. 

Wang, H. B., Ma, M. G., Xie, Y. M., Wang, X. F., and Wang, J.: Parameter inversion estimation in photosynthetic models: Impact of different simulation methods, Photosynthetica, 52, 233–246, https://doi.org/10.1007/s11099-014-0027-8, 2014. 

Wunsch, A., Liesch, T., and Broda, S.: Groundwater level forecasting with artificial neural networks: a comparison of long short-term memory (LSTM), convolutional neural networks (CNNs), and non-linear autoregressive networks with exogenous input (NARX), Hydrol. Earth Syst. Sci., 25, 1671–1687, https://doi.org/10.5194/hess-25-1671-2021, 2021. 

Xu, C., McDowell, N. G., Fisher, R. A., Wei, L., Sevanto, S., Christoffersen, B. O., Weng, E., and Middleton, R. S.: Increasing impacts of extreme droughts on vegetation productivity under climate change, Nat. Clim. Change, 9, 948–953, https://doi.org/10.1038/s41558-019-0630-6, 2019. 

Yin, X. and Struik, P. C.: C3 and C4 photosynthesis models: An overview from the perspective of crop modelling, NJAS: Wageningen Journal of Life Sciences, 57, 27–38, https://doi.org/10.1016/j.njas.2009.07.001, 2009. 

Zhang, E., Liu, L., and Huang, L.: Automatically delineating the calving front of Jakobshavn Isbræ from multitemporal TerraSAR-X images: a deep learning approach, The Cryosphere, 13, 1729–1741, https://doi.org/10.5194/tc-13-1729-2019, 2019. 

Zhang, X.-Y., Huang, Z., Su, X., Siu, A., Song, Y., Zhang, D., and Fang, Q.: Machine learning models for net photosynthetic rate prediction using poplar leaf phenotype data., PLoS One, 15, e0228645, https://doi.org/10.1371/journal.pone.0228645, 2020. 

Zhang, Z., Xin, Q., and Li, W.: Machine learning-based modeling of vegetation leaf area index and gross primary productivity across North America and comparison with a process-based model, J. Adv. Model. Earth Sy., 13, e2021MS002802, https://doi.org/10.1029/2021MS002802, 2021. 

Zhi, W., Feng, D., Tsai, W.-P., Sterle, G., Harpold, A., Shen, C., and Li, L.: From hydrometeorology to river water quality: Can a deep learning model predict dissolved oxygen at the continental scale?, Environ. Sci. Technol., 55, 2357–2368, https://doi.org/10.1021/acs.est.0c06783, 2021. 

Zhi, W., Ouyang, W., Shen, C., and Li, L.: Temperature outweighs light and flow as the predominant driver of dissolved oxygen in US rivers, Nat. Water, 1, 249–260, https://doi.org/10.1038/s44221-023-00038-z, 2023. 

Zhu, F., Li, X., Qin, J., Yang, K., Cuo, L., Tang, W., and Shen, C.: Integration of multisource data to estimate downward longwave radiation based on deep neural networks, IEEE T. Geosci. Remote Sens., 1–15, https://doi.org/10.1109/TGRS.2021.3094321, 2021. 

Download
Short summary
Photosynthesis is critical for life and has been affected by the changing climate. Many parameters come into play while modeling, but traditional calibration approaches face many issues. Our framework trains coupled neural networks to provide parameters to a photosynthesis model. Using big data, we independently found parameter values that were correlated with those in the literature while giving higher correlation and reduced biases in photosynthesis rates.
Altmetrics
Final-revised paper
Preprint