Next Article in Journal
Monitoring Ash Dieback in Europe—An Unrevealed Perspective for Remote Sensing?
Next Article in Special Issue
Estimation of Net Ecosystem Productivity on the Tibetan Plateau Grassland from 1982 to 2018 Based on Random Forest Model
Previous Article in Journal
Linear and Nonlinear Land Use Regression Approach for Modelling PM2.5 Concentration in Ulaanbaatar, Mongolia during Peak Hours
Previous Article in Special Issue
NPP Variability Associated with Natural and Anthropogenic Factors in the Tropic of Cancer Transect, China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Technical Note

A Radiation-Regulated Dynamic Maximum Light Use Efficiency for Improving Gross Primary Productivity Estimation

1
College of Water Sciences, Beijing Normal University, Beijing 100875, China
2
State Key Laboratory of Remote Sensing Science, Faculty of Geographical Science, Beijing Normal University, Beijing 100875, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(5), 1176; https://doi.org/10.3390/rs15051176
Submission received: 19 December 2022 / Revised: 7 February 2023 / Accepted: 20 February 2023 / Published: 21 February 2023

Abstract

:
The light use efficiency (LUE) model has been widely used in regional and global terrestrial gross primary productivity (GPP) estimation due to its simple structure, few input parameters, and particular theoretical basis. As a key input parameter of the LUE model, the maximum LUE (Ɛmax) is crucial for the accurate estimation of GPP and to the interpretability of the LUE model. Currently, most studies have assumed Ɛmax as a universal constant or constants depending on vegetation type, which means that the spatiotemporal dynamics of Ɛmax were ignored, leading to obvious uncertainties in LUE-based GPP estimation. Using quality-screened daily data from the FLUXNET 2015 dataset, this paper proposed a photosynthetically active radiation (PAR)-regulated dynamic Ɛmax (PAR-Ɛmax, corresponding model named PAR-LUE) by considering the nonlinear response of vegetation photosynthesis to solar radiation. The PAR-LUE was compared with static Ɛmax-based (MODIS and EC-LUE) and spatial dynamics Ɛmax-based (D-VPM) models at 171 flux sites. Validation results showed that (1) R2 and RMSE between PAR-LUE GPP and observed GPP were 0.65 (0.44) and 2.55 (1.82) g C m−2 MJ−1 d−1 at the 8-day (annual) scale, respectively; (2) GPP estimation accuracy of PAR-LUE was higher than that of other LUE-based models (MODIS, EC-LUE, and D-VPM), specifically, R2 increased by 29.41%, 2.33%, and 12.82%, and RMSE decreased by 0.36, 0.14, and 0.34 g C m−2 MJ−1 d−1 at the annual scale; and (3) specifically, compared to the static Ɛmax-based model (MODIS and EC-LUE), PAR-LUE effectively relieved the underestimation of high GPP. Overall, the newly developed PAR-Ɛmax provided an estimation method utilizing a spatiotemporal dynamic Ɛmax, which effectively reduced the uncertainty of GPP estimation and provided a new option for the optimization of Ɛmax in the LUE model.

1. Introduction

As the largest component of the terrestrial carbon cycle [1,2,3], the accurate estimation of terrestrial gross primary productivity (GPP) is vital for understanding global carbon cycle processes, climate changes, and ecosystem services [1,4,5,6]. To date, GPP estimation is difficult, as no direct measures are available at regional and global scales. The eddy covariance flux tower provides a direct measure of carbon dioxide (CO2) between the land surface and the atmosphere that can be used for indirect GPP estimation [7]. Flux-based GPP has been widely used as reference data for calibrating and validating GPP models [8,9,10,11,12]. With the development of GPP estimation theories and methods in recent decades, researchers have developed many remote sensing models for GPP estimation [3,13,14].
Light use efficiency (LUE) models have been widely used to estimate terrestrial GPP at regional and global scales due to their theoretical basis, few parameters, and high practicality [15]. Since Monteith [16] proposed the concept of LUE, following the theoretical basis of LUE [17], researchers have optimized the input parameters for calculating LUE and developed dozens of LUE models [2]. As a key parameter of the LUE model, the maximum LUE (LUEmax, same as Ɛmax) is crucial for the accurate estimation of GPP and for the interpretability of the LUE model. The Ɛmax in existing LUE models can be roughly divided into three main categories: (1) global constant Ɛmax, such as the constant Ɛmax used in C-Fix [18] and EC-LUE [12]; (2) constant Ɛmax for each type, such as the constant Ɛmax varying from vegetation types in MOD17 [17] and VPRM [19], the constant Ɛmax for C3 and C4 in VPM [10,20,21] and TEC [22], the constant Ɛmax for sunlit and shaded leaves in TL-LUE [23] and DTEC [24], and the constant Ɛmax for different phenological stages in TS-LUE [25]; and (3) dynamic Ɛmax, such as the cloudiness index-regulated dynamic Ɛmax in CFlux [26], CI-LUE [27], and CI-EF [28], and the spatial dynamic Ɛmax based on the enhanced vegetation index (EVI) and visible albedo [29].
With the development of LUE models, an increasing number of researchers have considered Ɛmax as a dynamic value rather than a constant. Dynamic Ɛmax is more consistent with vegetation physiology, and studies have proven that in their own study area, dynamic Ɛmax performed better than static Ɛmax in GPP estimation [30,31]. Although researchers have proposed methods to estimate dynamic Ɛmax, the nonlinear response of vegetation photosynthesis to solar radiation variation was rarely considered in these studies. As the energy source of vegetation photosynthesis, solar radiation variation directly regulated the vegetation Ɛmax. Most LUE models implied a linear relationship between GPP and PAR, which apparently ignored the saturation of vegetation photosynthesis to solar radiation under high solar radiation. In fact, vegetation photosynthesis varies with the dynamics of solar radiation [32]. For the low radiation situation, especially when vegetation photosynthesis is limited only by radiation, photosynthesis would increase linearly with increasing radiation. For the high radiation situation (plentiful radiation), photosynthesis would become radiation-saturated and no longer respond to the changes in radiation supply [32,33,34,35,36].
Considering the nonlinear response of vegetation photosynthesis to solar radiation [37], this paper proposed a photosynthetically active radiation (PAR)-regulated dynamic Ɛmax (PAR-Ɛmax, corresponding model named PAR-LUE) by using the quality-screened daily GPP and PAR data from the FLUXNET 2015 dataset. The performances of PAR-LUE in GPP estimation were evaluated based on the observed GPP and other LUE-based GPP estimation results.

2. Data and Preprocessing

2.1. FLUXNET Data

The FLUXNET 2015 dataset (https://fluxnet.fluxdata.org/data/FLUXNET2015-dataset/ (accessed on 1 August 2022)) includes multiple temporal scales (e.g., half-hourly, hourly, daily, weekly, monthly, and yearly) of observations, which contains the flux data of carbon, water, and energy collected from 212 sites around the globe. Data were quality-controlled and processed using uniform methods to improve the consistency and intercomparability across sites [38,39]. In this paper, a total of 171 sites (1104 site-years) with high quality data (“NEE_QC” > 0.75) were selected from the FLUXNET 2015 dataset (Figure 1), and the variables of the daily “GPP_NT_VUT_MEAN” and ” SW_IN_F” were selected as the reference daily GPP (g C m−2 d−1) and shortwave radiation (SR; W m−2). In addition, the daily PAR was calculated using the site observed shortwave radiation according to the empirical formula (i.e., PAR = 0.45 × SR × 0.0864; MJ m−2). More details about the FLUXNET 2015 dataset can be found in Pastorello, Trotta [38].

2.2. MODIS Data

Daily MODIS surface reflectance products (MCD43A4 Version 6) and 8-day composite MODIS GPP products (MOD17A2H Version 6) with 500 m resolution were used in this study. The MCD43A4 and MOD17A2H for each carbon flux site were downloaded from the NASA MODIS/VIIRS Land Product Subsets (https://modis.ornl.gov/globalsubset/ (accessed on 1 August 2022)). All MODIS data were filtered according to their quality label. MOD17A2H was used as one of the comparison data to evaluate the performance of PAR-LUE in GPP estimation, and MCD43A4 was used to calculate visible albedo, EVI, and land surface water index (LSWI). Albedo, EVI, and LSWI were calculated as follows:
Albedovisible = 0.331RRed + 0.424RBlue + 0.246RGreen
EVI = 2.5   ×   R NIR     R Red R NIR + 6 R Red   7.5 R Blue + 1
LSWI = R NIR   R SWIR R NIR   +   R SWIR
where RBlue, RRed, RNIR, and RSWIR are the reflectances of the blue, red, near infrared (NIR), and shortwave infrared (SWIR) bands, respectively.

3. Methods

3.1. PAR-LUE

The common structure of the LUE model can be formulated as follows:
GPP = PAR × FPAR × Ɛmax × f (Ts) × f (Ws)
where PAR is photosynthetically active radiation, FPAR is the fraction of absorbed PAR, Ɛmax is maximal light use efficiency, and f (Ts) and f (Ws) are the scaled environmental stress indices of temperature and water on LUE, respectively.
Referring to the existing research results [10,21], FPAR, f (Ts), and f (Ws) were calculated as follows:
FPAR = ( EVI     0.1 )   ×   1.25
T S = ( T     T min ) ( T     T max ) ( T     T min ) ( T     T max )     ( T     T opt ) 2  
W S = 1 + LSWI 1 + LSWI max
where T is the daily temperature; Tmin, Tmax, and Topt are 0 °C, 40 °C, and 20 °C, respectively; and LSWImax is the maximal LSWI in the growing season. Here, the growing season is defined according to the date of 75 days before and after the date of maximal EVI (i.e., [date_EVImax–75, date_EVImax + 75]).
In the PAR-LUE model, we proposed a PAR-based method to calculate dynamic Ɛmax (i.e., PAR-Ɛmax, and PAR-Ɛmax = f (PAR)). Considering the nonlinear response of vegetation photosynthesis to PAR, we developed the PAR-LUE model based on two hypothesises. First, under the ideal condition that vegetation was only related to the PAR and unconstrained by other biotic and abiotic conditions (i.e., FPAR, f (Ts), and f (Ws) are equal to 1), GPPi can be represented as the product of PAR and Ɛmax (8). Under the optimal temperature and water conditions (i.e., T = Topt, LSWI = LSWImax), the f (Ts) and f (Ws) are equal to 1. For the FPAR, the maximum value of measured [40] and remotely sensed FPAR (e.g., for the formula (5), FPAR = 1 when EVI ≥ 0.9) are close or equal to 1. Second, the maximum GPP (GPPmax) under different levels of PAR meets the ideal conditions (i.e., GPPi = GPPmax).
Using the quality-screened daily GPP and PAR data from the FLUXNET 2015 dataset, the maximum value of GPP corresponding to PAR (PAR was sampled with a step of 1 MJ m−2) was sampled within the subrange of 1 ± 0.25 MJ m−2. Then, the sampled PAR and GPPmax were fitted using cubic polynomial (9). Finally, combining Formulas (8) and (9), PAR-Ɛmax was calculated with Formula (10), and the corresponding model was named PAR-LUE.
GPPi = PAR × Ɛmax = PAR × f (PAR)
GPPmax = a × PAR3 + b × PAR2 + c × PAR
PAR-Ɛmax= a × PAR2 + b × PAR + c
where GPPi is the GPP under ideal conditions, and a, b, and c are fitting parameters of the cubic polynomial.
As shown in Figure 2, the orange fit curve indicated the vegetation GPPmax under different PAR levels, and the points below the fit curve indicated the true GPP constrained by vegetation physiology and environmental factors. It is important to note that PAR-Ɛmax is the Ɛmax of all vegetation in the ideal condition. For the differences in Ɛmax among vegetation types, it is expected that the EVI-based FPAR can regulate those differences under the framework of the LUE model.

3.2. Reference LUE Model

In this paper, static Ɛmax- and spatial dynamic Ɛmax-based LUE models were built as reference models (named EC-LUE and D-VPM, respectively) to examine the performances of the newly developed PAR-Ɛmax-based model (PAR-LUE) under the same LUE model framework. Among those models, only Ɛmax is different. Specifically, the static Ɛmax is 2.14 g C m−2 MJ−1 [12], while the spatial dynamic Ɛmax (named RS-Ɛmax) is calculated according to the EVI and albedo [29]:
{ ε max = exp ( 1.428 MaxE     6.295 MinVa     1.211 ) ,   MaxE   >   0.07 ε max = 0 , MaxE     0.07
where MaxE and MinVa are the maximal EVI and minimal visible albedo in the growing season, respectively.

3.3. Accuracy Evaluation

Two criteria were used here to evaluate model performance, including the determination coefficient (R2) and root mean square error (RMSE). In addition to the validation with observed GPP from FLUXNET sites, the performances of the PAR-LUE model in GPP estimation were compared with that of the MOD17 algorithm, EC-LUE and D-VPM from multiple dimensions, such as overall accuracy (8-day and annual), accuracy of each vegetation type and the seasonal dynamics in typical sites.

4. Results

4.1. Comparison of Different Ɛmax

Different PAR-Ɛmax calculated usingthe different GPPmax sample percentiles have similar performances in GPP estimation (Table 1). With the decrease in the GPPmax sample percentile, the R2 between the PAR-LUE-estimated GPP and observed GPP slightly increased, while the RMSE obviously increased (hereafter, PAR-Ɛmax was calculated according to GPPmax). At all selected flux sites, the spatiotemporal dynamic ranges of daily PAR-Ɛmax and RS-Ɛmax were 1.86–3.85 g C m−2 MJ−1 and 0.73–4.39 g C m−2 MJ−1, respectively (Figure 3). Obviously, the variation range of RS-Ɛmax was larger than that of PAR-Ɛmax, and both dynamic Ɛmax contained a constant value of 2.14 g C m−2 MJ−1.
PAR-Ɛmax showed significant seasonal dynamics, which presented as a “U” shaped variation in a natural year (Figure 4). The seasonal trends of PAR-Ɛmax were relatively similar at the 10 typical vegetation sites, and their annual minimal PAR-Ɛmax values were close to 2.14 g C m−2 MJ−1. However, although RS-Ɛmax showed spatial and interannual variations, there was no seasonal dynamic.

4.2. Comparison of GPP Estimation

The overall estimation accuracy of PAR-LUE GPP was better than that of MODIS GPP, EC-LUE GPP, and D-VPM GPP (Figure 5). Compared with MODIS GPP, EC-LUE GPP, and D-VPM GPP at the 8-day scale, the R2 between PAR-LUE GPP and observed GPP increased by 12.07%, 1.56%, and 8.33%, and the RMSE decreased by 0.18, 0.09, and 0.41 g C m−2 MJ−1 d−1, respectively. At the annual scale, R2 increased by 29.41%, 2.33%, and 12.82%, and the RMSE decreased by 0.36, 0.14, and 0.34 g C m−2 MJ−1 d−1, respectively. Although the GPP estimation accuracies of PAR-LUE and EC-LUE were closer in R2 and RMSE, EC-LUE obviously underestimated the high GPP. Overall, PAR-LUE showed better performances than the reference LUE models in GPP estimation, especially in reducing the underestimation of high GPP. From the performances of LUE models in different vegetation types (Figure 6), the GPP estimation accuracy of PAR-LUE was generally comparable to that of EC-LUE and D-VPM in most types.
The comparison at typical sites in the Northern Hemisphere showed that PAR-LUE, EC-LUE, and D-VPM were all in good agreement with the observed GPP in characterizing seasonal dynamics (Figure 7). In the majority of vegetation types, the PAR-LUE GPP was in better agreement with the observed GPP, which is evidenced by a smaller RMSE and a closer regression slope to 1. For example, at the evergreen coniferous forest site (Figure 7c), although the GPP estimated using the three LUE models showed very high agreement with the observed GPP (R2 ≥ 0.93), the PAR-LUE-estimated GPP was closer to the observed GPP throughout the growing season (smaller RMSE and closer slope to 1); at the closed shrub site (Figure 7g), the PAR-LUE GPP was closer to the observed GPP during the rising and falling phases of the growing season.

5. Discussion

As the direct energy source for vegetation photosynthesis, solar radiation directly determines the light use efficiency of vegetation. Therefore, Ɛmax should have corresponding spatiotemporal dynamics to the background of spatiotemporal solar radiation variation. The spatiotemporal dynamics of Ɛmax were the result of the long-term adaptation of vegetation to variations in solar radiation (especially seasonal variations), and vegetation usually has different Ɛmax values under different radiation conditions. The seasonal dynamics of PAR-Ɛmax presented in this paper show a “U”-shaped trend, with a large value in spring and autumn and a small value in summer. During vegetation green-up, vegetation has a large Ɛmax to make full use of the limited PAR and thus promote vegetation growth and development. However, during the peak growing season, solar radiation is more abundant, and vegetation photosynthesis tends to be saturated (high solar radiation will even reduce photosynthesis) and has a relatively small Ɛmax. There were similar explanations in the studies of Chapin and Matson [32] and Mõttus and Sulev [34]. The study of Propastin and Ibrom [41] in a tropical rainforest found that the LUE model would overestimate vegetation GPP under high radiation conditions if the saturation effect of vegetation photosynthesis on solar radiation is not considered. In addition, some studies showed that Ɛmax was different in clear and cloudy skies [37], as well as in sunlit and shaded leaves [31], which partly explains the influence of PAR on Ɛmax. Some cloudiness indices that regulate Ɛmax used in CFlux [26], CI-LUE [27], and CI-EF [28] can also be considered a sort of radiation-regulated Ɛmax because the cloudiness index was calculated based on PAR [42].
In this paper, a cubic polynomial function was used to calculate PAR-Ɛmax, which is simple to calculate and easy to fit. The fitted curve can effectively characterize the relationship between PAR and GPP in the actual range of PAR variation, and the shape of the fitted curve is consistent with existing studies [32,33,34,36]. The Ɛmax can be estimated based on different data and methods, and there are differences in their physical meanings [43]. Some studies estimated Ɛmax based on the flux data observed during the peak growing season [44,45,46], which obtained the specific Ɛmax under radiation saturation. In terms of the full growing season, Ɛmax of the peak growing season is only a special case of its seasonal dynamic changes. For example, the specific constant Ɛmax used in the EC-LUE model was comparable to the seasonal minimum of PAR-Ɛmax (Figure 3). In the comparison of Ɛmax defined by different studies, extra attention needs to be paid to their essential meanings. For example, the study of Zhang and Xiao [43] indicated that the daily Ɛmax exhibits less variation across biome types and seasons, which contradicts the spatiotemporal dynamics of PAR-Ɛmax. It is important to note that the Ɛmax defined by Zhang and Xiao [43] contains the FPAR, while the FPAR in PAR-Ɛmax is assumed to be 1. In terms of seasonal trends, PAR-Ɛmax was consistent with the reference LUE in the study of Zhang and Xiao [43].
Dynamic Ɛmax-based PAR-LUE performed better in GPP estimation than that of constant Ɛmax-based MODIS GPP and EC-LUE. In terms of GPP estimation accuracy alone (R2 and RMSE), EC-LUE was comparable to PAR-LUE in GPP estimation. However, the PAR-LUE mitigated the underestimation of high GPP, which is a nonnegligible contribution to an accurate estimate of total annual GPP. The constant Ɛmax in the EC-LUE model, as a special case of dynamic Ɛmax, is one of the reasons for its comparable ability to estimate GPP (especially R2) with the PAR-LUE model. On the one hand, the GPP values at the beginning and end of the growing season were relatively small, so only a minor difference existed between the GPP estimated by the constant and dynamic Ɛmax. For example, in the comparison of the seasonal variation in GPP at typical sites in the Northern Hemisphere (Figure 7), EC-LUE and PAR-LUE showed similar R2 values, but the seasonal dynamics of PAR-LUE GPP were closer to the observed GPP. On the other hand, the accuracy of other input parameters in the LUE model may inhibit the ability of dynamic Ɛmax to improve the accuracy of GPP estimation. In addition, compared with the spatial dynamics of Ɛmax-based D-VPM, PAR-LUE has two advantages in addition to the improvement in precision. First, PAR drives vegetation photosynthesis more directly than albedo, the Ɛmax constructed by PAR is more theoretical than that of albedo, and PAR-Ɛmax has both spatial and temporal dynamic characteristics. Second, the estimation of spatial dynamic RS- Ɛmax requires remote sensing data in a whole growing season (it requires the maximum EVI and minimum albedo of the whole growing season), which limits its application in near-real-time GPP estimation, while PAR-Ɛmax is not affected by this.
The accuracy validation results showed the reasonableness and reliability of PAR-Ɛmax. However, PAR-LUE still has room for improvement. First, the accurate estimation of PAR-Ɛmax requires a large number of high-quality observations. Limited by site representativeness and data quality, PAR-Ɛmax may have some errors in the specific values. A large number of high-quality observations will help to improve the performance of the PAR-LUE model. Second, PAR-Ɛmax characterizes the Ɛmax of all vegetation, and it was thought that the EVI-based FPAR could constrain the differences in PAR-Ɛmax across vegetation types in the PAR-LUE model. However, from the performances of PAR-LUE in estimating vegetation GPP (Figure 5, Figure 6 and Figure 7), FPAR was able to partly characterize differences in vegetation types, but its ability to constrain PAR-Ɛmax was still limited. Obviously, the PAR-Ɛmax of different vegetation types can be obtained based on vegetation type data, but it will be limited by the quality of vegetation type data, and it is difficult to characterize the spatially continuous variation in terrestrial vegetation with vegetation type data. It is hoped that further research can introduce a factor that can characterize the spatially continuous variation in vegetation photosynthetic capacity in PAR-LUE to improve the theory of the PAR-LUE model and the accuracy of GPP estimation. Finally, the effect of other input parameters of the LUE model on the accuracy of GPP estimation needs to be further analyzed to reduce the interference from other input parameters on the contribution of Ɛmax parameter optimization to the improvement of GPP estimation.

6. Conclusions

Considering the nonlinear response of vegetation photosynthesis to solar radiation, we proposed a new Ɛmax with spatiotemporal dynamics (PAR-Ɛmax, model denoted PAR-LUE), with using the daily PAR and GPP data observed from the flux tower, based on the assumption that GPP is only determined by PAR and Ɛmax under ideal conditions. Flux data-based validation results showed that the accuracy of PAR-LUE-estimated GPP was better than that of constant and spatially varied Ɛmax-based models. The PAR-LUE was suitable for remote sensing based GPP estimation of most vegetation types at regional and global scales. Overall, the newly developed PAR-Ɛmax provided an estimation method of spatiotemporal dynamic Ɛmax, which effectively reduced the uncertainty of GPP estimation and provided a new option for the optimization and development of dynamic Ɛmax in the LUE model.

Author Contributions

Conceptualization, Z.X., C.Z., W.Z. and Y.H.F.; methodology, Z.X. and C.Z.; writing—original draft preparation, Z.X. and C.Z.; writing—review and editing, Z.X., C.Z., H.Z., W.Z. and Y.H.F.; visualization, C.Z. and H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This study is funded by the National Key Research and Development Program of China (Grant No. 2020YFA0608504) and the Second Tibetan Plateau Scientific Expedition and Research Program (STEP) (Grant No. 2019QZKK0606).

Data Availability Statement

Data is available upon request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Beer, C.; Reichstein, M.; Tomelleri, E.; Ciais, P.; Jung, M.; Carvalhais, N.; Rödenbeck, C.; Arain, M.A.; Baldocchi, D.; Bonan, G.B.; et al. Terrestrial gross carbon dioxide uptake: Global distribution and covariation with climate. Science 2010, 329, 834–838. [Google Scholar] [CrossRef] [Green Version]
  2. Pei, Y.; Dong, J.; Zhang, Y.; Yuan, W.; Doughty, R.; Yang, J.; Zhou, D.; Zhang, L.; Xiao, X. Evolution of light use efficiency models: Improvement, uncertainties, and implications. Agric. For. Meteorol. 2022, 317, 108905. [Google Scholar] [CrossRef]
  3. Ryu, Y.; Berry, J.A.; Baldocchi, D.D. What is global photosynthesis? History, uncertainties and opportunities. Remote Sens. Environ. 2019, 223, 95–114. [Google Scholar] [CrossRef]
  4. Anav, A.; Friedlingstein, P.; Beer, C.; Ciais, P.; Harper, A.; Jones, C.; Murray-Tortarolo, G.; Papale, D.; Parazoo, N.C.; Peylin, P.; et al. Spatiotemporal patterns of terrestrial gross primary production: A review. Rev. Geophys. 2015, 53, 785–818. [Google Scholar] [CrossRef] [Green Version]
  5. Xia, J.; Niu, S.; Ciais, P.; Janssens, I.A.; Chen, J.; Ammann, C.; Arain, A.; Blanken, P.D.; Cescatti, A.; Bonal, D.; et al. Joint control of terrestrial gross primary productivity by plant phenology and physiology. Proc. Natl. Acad. Sci. USA 2015, 112, 2788. [Google Scholar] [CrossRef] [Green Version]
  6. Zhang, Y.; Kong, D.; Gan, R.; Chiew, F.H.S.; McVicar, T.R.; Zhang, Q.; Yang, Y. Coupled estimation of 500 m and 8-day resolution global evapotranspiration and gross primary production in 2002–2017. Remote Sens. Environ. 2019, 222, 165–182. [Google Scholar] [CrossRef]
  7. Aubinet, M.; Vesala, T.; Papale, D. Eddy Covariance: A Practical Guide to Measurement and Data Analysis; Springer Atmospheric Sciences; Springer: Dordrecht, The Netherlands, 2012; p. 438. [Google Scholar]
  8. Baldocchi, D. Measuring fluxes of trace gases and energy between ecosystems and the atmosphere—The state and future of the eddy covariance method. Glob. Change Biol. 2014, 20, 3600–3609. [Google Scholar] [CrossRef]
  9. Jung, M.; Schwalm, C.; Migliavacca, M.; Walther, S.; Camps-Valls, G.; Koirala, S.; Anthoni, P.; Besnard, S.; Bodesheim, P.; Carvalhais, N.; et al. Scaling carbon fluxes from eddy covariance sites to globe: Synthesis and evaluation of the FLUXCOM approach. Biogeosciences 2020, 17, 1343–1365. [Google Scholar] [CrossRef] [Green Version]
  10. Zhang, Y.; Xiao, X.; Wu, X.; Zhou, S.; Zhang, G.; Qin, Y.; Dong, J. A global moderate resolution dataset of gross primary production of vegetation for 2000–2016. Sci. Data 2017, 4, 170165. [Google Scholar] [CrossRef] [Green Version]
  11. Jiang, C.; Ryu, Y. Multi-scale evaluation of global gross primary productivity and evapotranspiration products derived from Breathing Earth System Simulator (BESS). Remote Sens. Environ. 2016, 186, 528–547. [Google Scholar] [CrossRef]
  12. Yuan, W.P.; Liu, S.G.; Zhou, G.S.; Zhou, G.Y.; Tieszen, L.L.; Baldocchi, D.; Bernhofer, C.; Gholz, H.; Goldstein, A.H.; Goulden, M.L.; et al. Deriving a light use efficiency model from eddy covariance flux data for predicting daily gross primary production across biomes. Agric. For. Meteorol. 2007, 143, 189–207. [Google Scholar] [CrossRef] [Green Version]
  13. Xiao, J.; Chevallier, F.; Gomez, C.; Guanter, L.; Hicke, J.A.; Huete, A.R.; Ichii, K.; Ni, W.; Pang, Y.; Rahman, A.F.; et al. Remote sensing of the terrestrial carbon cycle: A review of advances over 50 years. Remote Sens. Environ. 2019, 233, 111383. [Google Scholar] [CrossRef]
  14. Sun, Z.; Wang, X.; Zhang, X.; Tani, H.; Guo, E.; Yin, S.; Zhang, T. Evaluating and comparing remote sensing terrestrial GPP models for their response to climate variability and CO2 trends. Sci. Total Environ. 2019, 668, 696–713. [Google Scholar] [CrossRef]
  15. Running, S.W.; Thornton, P.E.; Nemani, R.; Glassy, J.M. Global Terrestrial Gross and Net Primary Productivity from the Earth Observing System. In Methods in Ecosystem Science; Sala, O.E., Jackson, R.B., Mooney, H.A., Howarth, R.W., Eds.; Springer: New York, NY, USA, 2000; pp. 44–57. [Google Scholar]
  16. Monteith, J.L. Solar radiation and productivity in tropical ecosystems. J. Appl. Ecol. 1972, 9, 747–766. [Google Scholar] [CrossRef] [Green Version]
  17. Running, S.W.; Nemani, R.R.; Heinsch, F.A.; Zhao, M.S.; Reeves, M.; Hashimoto, H. A Continuous Satellite-Derived Measure of Global Terrestrial Primary Production. BioScience 2004, 54, 547–560. [Google Scholar] [CrossRef]
  18. Veroustraete, F.; Sabbe, H.; Eerens, H. Estimation of carbon mass fluxes over Europe using the C-Fix model and Euroflux data. Remote Sens. Environ. 2002, 83, 376–399. [Google Scholar] [CrossRef]
  19. Mahadevan, P.; Wofsy, S.C.; Matross, D.M.; Xiao, X.; Dunn, A.L.; Lin, J.C.; Gerbig, C.; Munger, J.W.; Chow, V.Y.; Gottlieb, E.W. A satellite-based biosphere parameterization for net ecosystem CO2 exchange: Vegetation Photosynthesis and Respiration Model (VPRM). Glob. Biogeochem. Cycles 2008, 22, GB2005. [Google Scholar] [CrossRef] [Green Version]
  20. Xiao, X.; Hollinger, D.; Aber, J.; Goltz, M.; Davidson, E.A.; Zhang, Q.; Moore, B. Satellite-based modeling of gross primary production in an evergreen needleleaf forest. Remote Sens. Environ. 2004, 89, 519–534. [Google Scholar] [CrossRef]
  21. Xiao, X.; Zhang, Q.; Braswell, B.; Urbanski, S.; Boles, S.; Wofsy, S.; Moore, B.; Ojima, D. Modeling gross primary production of temperate deciduous broadleaf forest using satellite images and climate data. Remote Sens. Environ. 2004, 91, 256–270. [Google Scholar] [CrossRef]
  22. Yan, H.; Wang, S.-Q.; Billesbach, D.; Oechel, W.; Bohrer, G.; Meyers, T.; Martin, T.A.; Matamala, R.; Phillips, R.P.; Rahman, F.; et al. Improved global simulations of gross primary product based on a new definition of water stress factor and a separate treatment of C3 and C4 plants. Ecol. Model. 2015, 297, 42–59. [Google Scholar] [CrossRef]
  23. He, M.; Ju, W.; Zhou, Y.; Chen, J.; He, H.; Wang, S.; Wang, H.; Guan, D.; Yan, J.; Li, Y.; et al. Development of a two-leaf light use efficiency model for improving the calculation of terrestrial gross primary productivity. Agric. For. Meteorol. 2013, 173, 28–39. [Google Scholar] [CrossRef]
  24. Yan, H.; Wang, S.Q.; Yu, K.L.; Wang, B.; Yu, Q.; Bohrer, G.; Billesbach, D.; Bracho, R.; Rahman, F.; Shugart, H.H. A Novel Diffuse Fraction-Based Two-Leaf Light Use Efficiency Model: An Application Quantifying Photosynthetic Seasonality across 20 AmeriFlux Flux Tower Sites. J. Adv. Model. Earth Syst. 2017, 9, 2317–2332. [Google Scholar] [CrossRef] [Green Version]
  25. Huang, L.; Lin, X.; Jiang, S.; Liu, M.; Jiang, Y.; Li, Z.-L.; Tang, R. A two-stage light use efficiency model for improving gross primary production estimation in agroecosystems. Environ. Res. Lett. 2022, 17, 104021. [Google Scholar] [CrossRef]
  26. King, D.A.; Turner, D.P.; Ritts, W.D. Parameterization of a diagnostic carbon cycle model for continental scale application. Remote Sens. Environ. 2011, 115, 1653–1664. [Google Scholar] [CrossRef]
  27. Wang, S.; Huang, K.; Yan, H.; Yan, H.; Zhou, L.; Wang, H.; Zhang, J.; Yan, J.; Zhao, L.; Wang, Y.; et al. Improving the light use efficiency model for simulating terrestrial vegetation gross primary production by the inclusion of diffuse radiation across ecosystems in China. Ecol. Complex. 2015, 23, 1–13. [Google Scholar] [CrossRef]
  28. De Almeida, C.T.; Delgado, R.C.; Galvao, L.S.; Ramos, M.C. Improvements of the MODIS Gross Primary Productivity model based on a comprehensive uncertainty assessment over the Brazilian Amazonia. ISPRS J. Photogramm. Remote Sens. 2018, 145, 268–283. [Google Scholar] [CrossRef]
  29. Wang, H.; Jia, G.; Fu, C.; Feng, J.; Zhao, T.; Ma, Z. Deriving maximal light use efficiency from coordinated flux measurements and satellite data for regional gross primary production modeling. Remote Sens. Environ. 2010, 114, 2248–2258. [Google Scholar] [CrossRef]
  30. Muramatsu, K.; Furumi, S.; Soyama, N.; Daigo, M. Estimating the seasonal maximum light use efficiency. In Land Surface Remote Sensing, II, Proceedings of the SPIE Asia-Pacific Remote Sensing, Beijing, China, 13–16 October 2014; Jackson, T.J., Chen, J.M., Gong, P., Liang, S., Eds.; SPIE: Bellingham, WA, USA, 2014. [Google Scholar]
  31. Lin, X.; Chen, B.; Chen, J.; Zhang, H.; Sun, S.; Xu, G.; Guo, L.; Ge, M.; Qu, J.; Li, L.; et al. Seasonal fluctuations of photosynthetic parameters for light use efficiency models and the impacts on gross primary production estimation. Agric. For. Meteorol. 2017, 236, 22–35. [Google Scholar] [CrossRef] [Green Version]
  32. Chapin, F.S.; Matson, P.A.; Vitousek, P.M. Carbon inputs to ecosystems. In Principles of Terrestrial Ecosystem Ecology; Chapin, F.S., Matson, P.A., Vitousek, P.M., Eds.; Springer: New York, NY, USA, 2011; pp. 123–156. [Google Scholar]
  33. Farquhar, G.D.; Von Caemmerer, S.; Berry, J.A. A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta 1980, 149, 78–90. [Google Scholar] [CrossRef] [Green Version]
  34. Mõttus, M.; Sulev, M.; Baret, F.; Lopez-Lozano, R.; Reinart, A. Photosynthetically Active Radiation: Measurement and Modeling. In Encyclopedia of Sustainability Science and Technology; Meyers, R.A., Ed.; Springer: New York, NY, USA, 2012; pp. 7902–7932. [Google Scholar]
  35. Law, B.; Falge, E.; Gu, L.; Baldocchi, D.; Bakwin, P.; Berbigier, P.; Davis, K.; Dolman, A.; Falk, M.; Fuentes, J.; et al. Environmental controls over carbon dioxide and water vapor exchange of terrestrial vegetation. Agric. For. Meteorol. 2002, 113, 97–120. [Google Scholar] [CrossRef] [Green Version]
  36. Zheng, T.; Chen, J.; He, L.; Arain, M.A.; Thomas, S.C.; Murphy, J.G.; Geddes, J.A.; Black, T.A. Inverting the maximum carboxylation rate (Vcmax) from the sunlit leaf photosynthesis rate derived from measured light response curves at tower flux sites. Agric. For. Meteorol. 2017, 236, 48–66. [Google Scholar] [CrossRef]
  37. Mercado, L.M.; Bellouin, N.; Sitch, S.; Boucher, O.; Huntingford, C.; Wild, M.; Cox, P.M. Impact of changes in diffuse radiation on the global land carbon sink. Nature 2009, 458, 1014–1017. [Google Scholar] [CrossRef] [Green Version]
  38. Pastorello, G.; Trotta, C.; Canfora, E.; Chu, H.; Christianson, D.; Cheah, Y.-W.; Poindexter, C.; Chen, J.; Elbashandy, A.; Humphrey, M.; et al. The FLUXNET 2015 dataset and the ONEFlux processing pipeline for eddy covariance data. Sci. Data 2020, 7, 225. [Google Scholar] [CrossRef]
  39. Reichstein, M.; Falge, E.; Baldocchi, D.; Papale, D.; Aubinet, M.; Berbigier, P.; Bernhofer, C.; Buchmann, N.; Gilmanov, T.; Granier, A.; et al. On the separation of net ecosystem exchange into assimilation and ecosystem respiration: Review and improved algorithm. Glob. Chang. Biol. 2005, 11, 1424–1439. [Google Scholar] [CrossRef]
  40. Li, W.; Fang, H. Estimation of direct, diffuse, and total FPARs from Landsat surface reflectance data and ground-based estimates over six FLUXNET sites. J. Geophys. Res. Biogeosci. 2015, 120, 96–112. [Google Scholar] [CrossRef]
  41. Propastin, P.; Ibrom, A.; Knohl, A.; Erasmi, S. Effects of canopy photosynthesis saturation on the estimation of gross primary productivity from MODIS data in a tropical forest. Remote Sens. Environ. 2012, 121, 252–260. [Google Scholar] [CrossRef]
  42. Turner, D.P.; Ritts, W.D.; Styles, J.M.; Yang, Z.; Cohen, W.B.; Law, B.E.; Thornton, P.E. A diagnostic carbon flux model to monitor the effects of disturbance and interannual variation in climate on regional NEP. Tellus B Chem. Phys. Meteorol. 2006, 58, 476–490. [Google Scholar] [CrossRef] [Green Version]
  43. Zhang, Y.; Xiao, X.; Wolf, S.; Wu, J.; Wu, X.; Gioli, B.; Wohlfahrt, G.; Cescatti, A.; van der Tol, C.; Zhou, S.; et al. Spatio-Temporal Convergence of Maximum Daily Light-Use Efficiency Based on Radiation Absorption by Canopy Chlorophyll. Geophys. Res. Lett. 2018, 45, 3508–3519. [Google Scholar] [CrossRef]
  44. Wang, J.; Xiao, X.; Wagle, P.; Ma, S.; Baldocchi, D.; Carrara, A.; Zhang, Y.; Dong, J.; Qin, Y. Canopy and climate controls of gross primary production of Mediterranean-type deciduous and evergreen oak savannas. Agric. For. Meteorol. 2016, 226–227, 132–147. [Google Scholar] [CrossRef] [Green Version]
  45. Xiao, X. Light absorption by leaf chlorophyll and maximum light use efficiency. IEEE Trans. Geosci. Remote Sens. 2006, 44, 1933–1935. [Google Scholar] [CrossRef]
  46. Wagle, P.; Xiao, X.; Torn, M.S.; Cook, D.R.; Matamala, R.; Fischer, M.L.; Jin, C.; Dong, J.; Biradar, C. Sensitivity of vegetation indices and gross primary production of tallgrass prairie to severe drought. Remote Sens. Environ. 2014, 152, 1–14. [Google Scholar] [CrossRef]
Figure 1. Spatial distribution of selected FLUXNET 2015 sites. (Background is the map of the MCD12Q1 in 2014; CRO: cropland (19 sites), CSH: closed shrub (2 sites), DBF: deciduous broadleaf forest (22 sites), DNF: deciduous needleleaf forests (0 site), EBF: evergreen broad-leaf forest (11 sites), ENF: evergreen needleleaf forest (44 sites), MF: mixed forest (8 sites), GRA: grassland (30 sites), OSH: open shrub (11 sites), SAV: savanna (7 sites), WSA: woody savanna (6 sites), WET: wetland (11 sites)).
Figure 1. Spatial distribution of selected FLUXNET 2015 sites. (Background is the map of the MCD12Q1 in 2014; CRO: cropland (19 sites), CSH: closed shrub (2 sites), DBF: deciduous broadleaf forest (22 sites), DNF: deciduous needleleaf forests (0 site), EBF: evergreen broad-leaf forest (11 sites), ENF: evergreen needleleaf forest (44 sites), MF: mixed forest (8 sites), GRA: grassland (30 sites), OSH: open shrub (11 sites), SAV: savanna (7 sites), WSA: woody savanna (6 sites), WET: wetland (11 sites)).
Remotesensing 15 01176 g001
Figure 2. Schematic of PAR-Ɛmax based on the relationship between daily observed GPP and PAR. (The color from purple to yellow indicates that the point density increases gradually. The orange fit curve indicates the vegetation GPPmax at different PAR levels, and the dashed curves from top to bottom are fitted curves at the 99th, 98th, 95th, and 90th percentiles of GPPmax. The fitted curve indicates the unconstrained ideal GPP, while the points below the curve indicate the true GPP constrained by vegetation physiology and environmental factors.)
Figure 2. Schematic of PAR-Ɛmax based on the relationship between daily observed GPP and PAR. (The color from purple to yellow indicates that the point density increases gradually. The orange fit curve indicates the vegetation GPPmax at different PAR levels, and the dashed curves from top to bottom are fitted curves at the 99th, 98th, 95th, and 90th percentiles of GPPmax. The fitted curve indicates the unconstrained ideal GPP, while the points below the curve indicate the true GPP constrained by vegetation physiology and environmental factors.)
Remotesensing 15 01176 g002
Figure 3. Variation range comparison of three daily Ɛmax at all selected flux sites. (The top and bottom edges of the box are the 75% and 25% quartiles, respectively; the short horizontal line and the small square in the middle of the box are the median and mean, respectively; the shape of the violin displays frequencies of values.)
Figure 3. Variation range comparison of three daily Ɛmax at all selected flux sites. (The top and bottom edges of the box are the 75% and 25% quartiles, respectively; the short horizontal line and the small square in the middle of the box are the median and mean, respectively; the shape of the violin displays frequencies of values.)
Remotesensing 15 01176 g003
Figure 4. Seasonal dynamics comparison of three Ɛmax at typical sites. Ten sites from different vegetation types were selected and shown in two adjacent years (2011–2012) to exhibit the differences in seasonal variation among the three Ɛmax.
Figure 4. Seasonal dynamics comparison of three Ɛmax at typical sites. Ten sites from different vegetation types were selected and shown in two adjacent years (2011–2012) to exhibit the differences in seasonal variation among the three Ɛmax.
Remotesensing 15 01176 g004
Figure 5. Accuracy comparison of GPP estimated from different LUE models. (ad) show the 8 daily results, and (eh) show the yearly results.
Figure 5. Accuracy comparison of GPP estimated from different LUE models. (ad) show the 8 daily results, and (eh) show the yearly results.
Remotesensing 15 01176 g005
Figure 6. Accuracy comparison of GPP estimated from different LUE models in different vegetation types. (a) Comparison of RMSE at the 8-day scale, (b) comparison of R2 at the 8-day scale, (c) comparison of RMSE at the yearly scale, and (d) comparison of R2 at the yearly scale.
Figure 6. Accuracy comparison of GPP estimated from different LUE models in different vegetation types. (a) Comparison of RMSE at the 8-day scale, (b) comparison of R2 at the 8-day scale, (c) comparison of RMSE at the yearly scale, and (d) comparison of R2 at the yearly scale.
Remotesensing 15 01176 g006
Figure 7. Comparison of GPP seasonal variation at typical sites. Eight site-year samples ((ah), Lat: latitude; Lon: longitude; see Figure 1 for vegetation types) with similar or concurrently higher R2 values were selected to exhibit the better performance of the PAR-LUE model in further reducing the RMSE.
Figure 7. Comparison of GPP seasonal variation at typical sites. Eight site-year samples ((ah), Lat: latitude; Lon: longitude; see Figure 1 for vegetation types) with similar or concurrently higher R2 values were selected to exhibit the better performance of the PAR-LUE model in further reducing the RMSE.
Remotesensing 15 01176 g007
Table 1. PAR-Ɛmax estimation coefficients and GPP estimation accuracy under different sampling percentiles.
Table 1. PAR-Ɛmax estimation coefficients and GPP estimation accuracy under different sampling percentiles.
PercentileabcR2RMSE
100th0.00030−0.123763.849510.612022.8979
99th−0.00039−0.067682.591940.610192.9812
98th0.00052−0.090872.520620.613433.1780
95th0.00073−0.090352.248670.614923.4993
90th0.00033−0.071561.913750.614633.7946
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Xie, Z.; Zhao, C.; Zhu, W.; Zhang, H.; Fu, Y.H. A Radiation-Regulated Dynamic Maximum Light Use Efficiency for Improving Gross Primary Productivity Estimation. Remote Sens. 2023, 15, 1176. https://doi.org/10.3390/rs15051176

AMA Style

Xie Z, Zhao C, Zhu W, Zhang H, Fu YH. A Radiation-Regulated Dynamic Maximum Light Use Efficiency for Improving Gross Primary Productivity Estimation. Remote Sensing. 2023; 15(5):1176. https://doi.org/10.3390/rs15051176

Chicago/Turabian Style

Xie, Zhiying, Cenliang Zhao, Wenquan Zhu, Hui Zhang, and Yongshuo H. Fu. 2023. "A Radiation-Regulated Dynamic Maximum Light Use Efficiency for Improving Gross Primary Productivity Estimation" Remote Sensing 15, no. 5: 1176. https://doi.org/10.3390/rs15051176

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop