Next Article in Journal
Comparison of a Bottom-Up GNSS Radio Occultation Method to Measure D- and E-Region Electron Densities with Ionosondes and FIRI
Previous Article in Journal
Drought-Related Spatiotemporal Cumulative and Time-Lag Effects on Terrestrial Vegetation across China
Previous Article in Special Issue
Dynamic of Fluorescence Emissions at O2A and O2B Telluric Absorption Bands in Forested Areas with Seasonal APAR and GPP Variations
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Improving the Estimation of Canopy Fluorescence Escape Probability in the Near-Infrared Band by Accounting for Soil Reflectance

1
Key Laboratory of Digital Earth Science, Aerospace Information Research Institute, Chinese Academy of Sciences, Beijing 100094, China
2
International Research Center of Big Data for Sustainable Development Goals, Beijing 100094, China
3
College of Resources and Environment, University of Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Remote Sens. 2023, 15(18), 4361; https://doi.org/10.3390/rs15184361
Submission received: 28 July 2023 / Revised: 25 August 2023 / Accepted: 2 September 2023 / Published: 5 September 2023

Abstract

:
Solar-induced chlorophyll fluorescence (SIF) has been found to be a useful indicator of vegetation’s gross primary productivity (GPP). However, the directional SIF observations obtained from a canopy only represent a portion of the total fluorescence emitted by all the leaf photosystems because of scattering and reabsorption effects inside the leaves and canopy. Hence, it is crucial to downscale the SIF from canopy level to leaf level by modeling fluorescence escape probability (fesc) for improved comprehension of the relationship between SIF and GPP. Most methods for estimating fesc rely on the assumption of a “black soil background,” ignoring soil reflectance and the effect of scattering between soils and leaves, which creates significant uncertainties for sparse canopies. In this study, we added a correction factor considering soil reflectance, which was modeled using the Gaussian process regression algorithm, to the semi-empirical NIRv/FAPAR model and obtained the improved fesc model accounting for soil reflectance (called the fesc_GPR-SR model), which is suitable for near-infrared SIF downscaling. The evaluation results using two simulation datasets from the Soil–Canopy–Observation of Photosynthesis and the Energy Balance (SCOPE) model and the Discrete Anisotropic Radiative Transfer (DART) model showed that the fesc_GPR-SR model outperformed the NIRv/FAPAR model, especially for sparse vegetation, with higher accuracy for estimating fesc (R2 = 0.954 and RMSE = 0.012 for SCOPE simulations; R2 = 0.982 and RMSE = 0.026 for DART simulations) compared with the NIRv/FAPAR model (R2 = 0.866 and RMSE = 0.100 for SCOPE simulations; R2 = 0.984 and RMSE = 0.070 for DART simulations). The evaluation results using in situ observation data from multi-species canopies also suggested that the leaf-level SIF calculated by the fesc_GPR-SR model tracked better with photosynthetic active radiation absorbed by green components (APARgreen) for sparse vegetation (R2 = 0.937, RMSE = 0.656 mW/m2/nm) compared with the NIRv/FAPAR model (R2 = 0.921, RMSE = 0.904 mW/m2/nm). The leaf-level SIF calculated by the fesc_GPR-SR model was less sensitive to observation angles and differences in canopy structure among multiple species. These results emphasize the significance of accounting for soil reflectance in the estimation of fesc and demonstrate that the fesc_GPR-SR model can contribute to further exploring the physiological mechanism between SIF and GPP.

1. Introduction

Solar-induced chlorophyll fluorescence (SIF) is a phenomenon in which chlorophyll molecules in plants emit light in response to natural light. It is a byproduct of the light reaction process of photosynthesis [1,2] and has been proven to serve as a proxy for gross primary productivity (GPP) in numerous studies [2,3,4,5,6]. The light energy absorbed by photosynthetic pigments can be dissipated through three main pathways: photosynthesis, heat dissipation, and fluorescence [2,7]. SIF changes in sync with photochemical quenching and competes with heat dissipation under natural light conditions without stress, and it is closely tied to photosynthesis in the basic physiological and biochemical processes of plants [8,9]. As such, SIF provides a more accurate measurement of photosynthetic ecological changes and has a more direct relationship with GPP compared with vegetation indices based on reflectance [10,11].
There are currently many algorithms that can successfully retrieve SIF from ground-based and satellite-based observations [8,9,12,13,14]. However, the fluorescence photons observed at the canopy level only represent a fraction of the total SIF emission due to the effect of the radiative transfer process. After fluorescence photons are emitted by photosystems, they are absorbed and scattered by the leaf and canopy components and then escape to the canopy, where they are detected by sensors [2]. The red SIF is primarily affected by chlorophyll absorption within leaves, whereas the near-infrared SIF is primarily impacted by the scattering effect within the canopy [15]. The scattering and reabsorption effects result in different SIF–GPP relationships due to variations in canopy structure, such as leaf area, leaf orientation, and leaf clumping [16]. Furthermore, multiple scatterings inside the canopy during SIF transmission make the SIF observed from different angles vary, which demonstrates that canopy SIF is directional and not isotropic [17]. As a result, the canopy-level SIF differs from the leaf-level SIF and cannot be utilized to directly quantify changes in plant physiology [5,18,19].
To reduce the influence of the canopy structure and directional effect on the SIF–GPP relationship, it is necessary to downscale the SIF from canopy level to leaf level and obtain the leaf-level SIF which has a closer physiological coupling relationship with GPP. The fluorescence escape probability (fesc) represents the probability that the fluorescence emitted by photosystems escapes from the canopy and is an important bridge connecting the canopy SIF (SIFcanopy) and the total SIF emitted by photosystems (SIFtotal). Their relationship can be expressed as follows [20,21,22,23,24,25]:
S I F c a n o p y = S I F t o t a l × f e s c = P A R × F A P A R × Φ S I F × f e s c
where PAR is the incident photosynthetically active radiation; FAPAR is the fraction of photosynthetically active radiation absorbed by vegetation; ΦSIF is the fluorescence quantum yield. The weak absorption effect of leaves in the near-infrared band means that the fluorescence escape probability from photosystems to leaves can be approximated by the leaf albedo with a value around 1 [26], and the fesc in Equation (1) can be approximated as the fluorescence escape probability from leaves to the canopy and its unit is sr−1.
In recent years, a number of studies have concentrated on estimating fesc, in other words, downscaling SIF from the canopy to the leaf level; however, a common problem in most of these approaches is that they are based on spectrally invariant properties and the assumption of a “black soil background,” assuming that the soil is a black body with a reflectance of 0, which absorbs all external radiation signals, regardless of the reflective property of the soil and the effects of multiple scatterings among soils and leaves. Yang et al. [27] proposed that fesc could be accurately estimated using near-infrared reflectance, canopy interception (i0), and leaf albedo in the near-infrared band (ωN), assuming no soil reflectance (fesc = RefNIR/i0·ωN), which laid the foundation for the study of fesc estimation. Liu et al. [15] used this model to estimate fesc from canopy reflectance based on random forest regression, effectively avoiding the difficulties of estimating i0 and ωN in the model. Zeng et al. [28] further simplified the parameters in the model by developing the near-infrared reflectance of vegetation index (NIRv) to isolate the contribution of soil and vegetation to canopy reflectance and using FAPAR to approximate i0 in the denominator. Similarly, Liu et al. [26] derived fluorescence escape probability formulas for the red and near-infrared bands based on reflectance and FAPAR, respectively, and applied them to correct long-term ground observation data of SIF. Yang et al. [29] later showed that it was challenging to estimate FAPAR and fesc in the near-infrared band separately using canopy reflectance alone; however, the product of the two could be well replaced with a new vegetation index, the fluorescence-corrected vegetation index (FCVI), which was shown to be in good agreement with spectrally invariant properties.
Among the above-mentioned SIF-downscaling methods, the semi-empirical model proposed by Zeng et al. [28] is simple and accurate, facilitating the use of current in situ or remotely sensed NIRv and FAPAR datasets for fesc estimation. As a result, Zeng et al.’s model has been widely used and can be expressed as:
f e s c N I R v i 0 × ω N = N D V I × Re f N I R i 0 × ω N N I R v F A P A R
where fesc is a dimensionless quantity, different from fesc (unit: sr−1) in Equation (1). For ease of understanding, the fesc values in the following sections are all expressed as dimensionless quantities. The canopy interception, i0, is approximated by FAPAR. The leaf single scattering albedo in the near-infrared band (leaf reflectance + transmittance), ωN, is taken to be constant and equal to 1. Although the use of NIRv in the model is intended to eliminate the effect of soil reflectance, in practice, NIRv as a pure vegetation signal is controversial because soil reflectance impacts the calculation of the normalized difference vegetation index (NDVI). Additionally, the approximation of i0 with FAPAR relies on the assumption that soil reflectance is 0 [29]. Hence, the NIRv/FAPAR model proposed by Zeng et al. does not completely eliminate the impact of soil reflectance, which remains a key source of uncertainty in fesc estimation, especially for sparse vegetation. The influence of soil reflectance on fesc estimation can be seen in two aspects: its impact on the calculation of pure vegetation reflectance and on the scattering process between the canopy and soil.
To thoroughly correct the effect of soil reflectance, we aim to add a correction factor that includes soil reflectance to the simple NIRv/FAPAR model to increase the precision of fesc and leaf-level SIF estimation in the near-infrared band. First, we used the Soil–Canopy–Observation of Photosynthesis and the Energy Balance (SCOPE) model to simulate the training dataset. Then, we employed the Gaussian process regression (GPR) algorithm to model the correction factor, obtaining the improved fesc modeling method accounting for soil reflectance (fesc_GPR-SR model). Finally, we assessed the performance of the fesc_GPR-SR model using simulated data, including the SCOPE model and the Discrete Anisotropic Radiative Transfer (DART) model, combined with field-measured data.

2. Materials and Methods

2.1. Simulated Datasets

2.1.1. SCOPE Model Simulations

The SCOPE model [30] is a one-dimensional radiative transfer model that simulates the interaction between radiative transport, microscopic meteorological processes, spectral reflectance, SIF, and photosynthetic and hydrothermal fluxes in leaves and canopies. We used SCOPE v1.73 to simulate canopy-level SIF, leaf-level SIF, and canopy directional reflectance.
Leaf absorption is primarily influenced by chlorophyll concentration [31], while canopy scattering is primarily influenced by Sun–target–viewing geometries and structural parameters such as leaf area index (LAI) and leaf inclination distribution function (LIDF) [32]. We parameterized the model with a range of leaf chlorophyll a and b content (Cab), LAI, and five typical LIDFs (excluding Erectophile, which is uncommon) to cover the most common vegetation states. Since only the soil reflectance at 780 nm is used for model training, it has nothing to do with the shape of the soil spectra. It is only necessary to ensure that the soil spectra input of the SCOPE model covers the common soil reflectance range in the near-infrared band [33] (Figure 1). We selected small LAI values densely to represent sparse vegetation conditions. The details of the input variables for the SCOPE model are listed in Table 1. In total, 81,000 distinct samples were generated. To improve model training efficiency and reduce the processing time, we randomly selected 9000 samples for training (1500 simulated samples per soil spectral curve, 1500 × 6 = 9000), and another 9000 samples were selected for validation.

2.1.2. DART Model Simulations

The DART model [34] is a three-dimensional radiative transfer model that simulates the propagation of radiation across the whole optical domain, from the visible to the thermal infrared region, for natural landscapes such as forests, grasslands, and farmland, as well as for urban landscapes with topography and atmosphere. Recently, the FLUSPECT model was integrated into the DART model to simulate the radiative transfer of SIF within 3-D canopies [35]. In this study, the accuracy of the fesc was verified using DART v5.6.7, which simulated the SIF at both canopy and leaf levels, as well as the directional reflectance for fifty various viewing angles for maize canopies. The input variables for the DART model are listed in Table 2. To simulate sparse maize canopies, the value of LAI was set to 2. The soil spectrum from the DART model database was used (Figure S1). The specific simulated 3-D maize canopy scene is shown in Figure S2. Figure 2 shows the simulated multi-angle canopy SIF results in the near-infrared band (760 nm).

2.2. Field Dataset

2.2.1. Field-Measured Dataset

The in situ spectral data were collected from four field experiments carried out at three locations in 2016. The data were employed to assess the performance of the improved fesc estimation model in canopies with varying structures. The first and second spectral measurements of winter wheat were taken at the Xiaotangshan Farm (XTS) in Beijing on 8–9 April, 18 April, and 8 December, when the winter wheat was at the stages of jointing, booting, and tillering, respectively. The third experiment was conducted on 18 December at the Nanbin Farm (NBF) in Sanya and included vegetables and crops such as sweet potato, cotton, pumpkin, and maize. The fourth spectral measurement was taken on 18 December at the Sanya Remote Sensing Satellite Data Receiving Station (SYS) and included gold coin grass. The details of these four ground measurements are listed in Table 3. According to the visual judgment of canopies in Figure 3, the LIDF type in XTS was spherical and, in NBF and SYS, it was planophile. All the spectral measurements were performed using a custom-made Ocean Optics QE Pro spectrometer (Ocean Optic, Inc., Dunedin, FL, USA).

2.2.2. SIF Retrieval

The three-band fluorescence line discriminator (3FLD) method [36] is simple and reliable for data with 0.3 nm spectral resolution, according to Damm et al. [37] and Liu et al. [38]. This method was selected in this study for retrieving SIF. The formula for the 3FLD method is expressed as follows:
S I F c a n o p y = ( I l e f t × w l e f t + I r i g h t × w r i g h t ) × L i n I i n × ( L l e f t × w l e f t + L r i g h t × w r i g h t ) ( I l e f t × w l e f t + I r i g h t × w r i g h t ) I i n
where I is the incident solar radiance reaching the top of the canopy; L is the entire upwelling radiance; and the weight w, which is inversely proportional to the distance between the left- and right-side bands and the inner band, is used to average I and L outside the absorption band. The subscripts in, left, and right denote the bands that are within, to the left of, and to the right of the absorption band, respectively. As noted by Liu et al. [38] and Liu et al. [22], the wavelengths corresponding to the left, inner, and right shoulder of the absorption feature for the O2-A band are 752.92 nm, 760.72 nm, and 768.87 nm, respectively.

2.2.3. Estimation of APARgreen

Because only the green components of the canopy can carry out photosynthesis, the FAPAR for the entire canopy can be divided into the photosynthetic active green components (FAPARgreen) and the non-photosynthetic active components (FAPARnon-green) [39]. Liu et al. [40] proposed an in situ measurement approach of FAPARgreen in the low vegetation canopy using a digital camera and reference panel, which has been proven to be effective. Using this method, FAPARgreen can be calculated as follows:
F A P A R g r e e n = P A R i P A R r ( A P A R exp _ b + A P A R cov _ b ) P A R i
where PARi and PARr are the incident and reflected PAR calculated from the DN values of digital photographs. APARexp_b and APARcov_b are the PAR absorbed by the exposed background and the vegetation-covered background, respectively. Consequently, photosynthetically active radiation absorbed by green components (APARgreen) can be further deduced as:
A P A R g r e e n = P A R × F A P A R g r e e n

2.3. Correction Factor Accounting for Soil Reflectance

The NIRv/FAPAR model, proposed by Zeng et al. [28] (Equation (2)), is currently a popular method for downscaling SIF in the near-infrared band. However, this method is limited by the fact that it does not explicitly consider the impact of soil reflectance, which can result in errors in fesc estimation, particularly for sparse vegetation. Therefore, a correction factor is necessary to optimize the estimation of fesc in Zeng et al. [28] to reduce the impact of soil reflectance. The contribution of soil background to fesc estimation is not only related to its reflectance but also to vegetation coverage, which can be represented by the vegetation index. Hence, the correction factor can be expressed as a function of soil reflectance in the near-infrared band and vegetation index—f (Refsoil, VI). In this study, four vegetation indices (NDVI, simple ratio (SR), two-band enhanced vegetation index without the blue band (EVI2), and perpendicular vegetation index (PVI)) were selected to characterize vegetation structure (Table 4). The correction factor was added to the widely used NIRv/FAPAR model by multiplying and adding functions, resulting in two improved models:
f e s c = N I R v F A P A R × f 1 ( Ref soil , VI )
f e s c = N I R v F A P A R + f 2 ( Ref soil , VI )
where f1 (Refsoil, VI) and f2 (Refsoil, VI) are the two correction factors for multiplicative correction and additive correction, respectively.
To improve the robustness of fesc estimation, a machine learning method was employed to estimate the correction factor f (Refsoil, VI). A number of machine learning algorithms were tested for model training, including decision tree, random forest, support vector machine, and Gaussian process regression (GPR), and it was found that the exponential Gaussian process regression [45] was the best training model in this study (Figure S3), which can quantitatively determine the prediction uncertainty in a systematic manner. The model was trained using the 5-fold cross-validation method. The detailed GPR model parameters can be found in Table S1. According to Equations (6) and (7), the input parameters of the two machine learning models were Refsoil and VI, and the outputs were f e s c / N I R v F A P A R and f e s c N I R v F A P A R , respectively. The Gaussian regression algorithm was used to train two different correction factors, f1 (Refsoil, VI) and f2 (Refsoil, VI), in the models. All input and output parameters can be obtained from the SCOPE simulations.

3. Results

3.1. Performance of Different Correction Factors for fesc Estimation

As discussed in Section 2.3, vegetation indices should be included in the correction factor of the model. In this study, NDVI, SR, EVI2, and PVI were evaluated. The determination coefficient (R2), root mean square error (RMSE), and mean absolute error (MAE) were used to evaluate the models’ performance. As depicted in Figure 4, model f e s c / N I R v F A P A R performed significantly better than model f e s c N I R v F A P A R . However, changing the vegetation indices had little effect on the R2 values. By comparing other performance evaluation indices (Table 5), we found that NDVI performed best under all circumstances, including sparse vegetation. NDVI is the most extensively used of over 40 vegetation indices [46] and has been proven as a reliable representation of LAI and FVC [47,48]. Hence, NDVI was finally chosen as the vegetation index in the correction factor.
When the inputs of the model were the combination of NDVI and Refsoil and the output was f e s c / N I R v F A P A R , the model performance R2 reached 0.850, indicating that f1 (Refsoil, NDVI) can explain 85% of the remaining part of fesc from the NIRv/FAPAR model. Even for sparse canopies (LAI < 2), R2 reached 0.690. Therefore, the improved fesc model in the near-infrared band accounting for soil reflectance (referred to as the fesc_GPR-SR model) is defined as follows:
f e s c = N I R v F A P A R × f 1 ( Ref soil ,   NDVI )

3.2. Evaluation of the fesc_GPR-SR Model Using Simulated Data

3.2.1. Validation of the fesc_GPR-SR Model Using SCOPE Simulations

The performance for estimating fesc of the fesc_GPR-SR model was first assessed using the SCOPE simulations. The comparison of fesc calculated by the NIRv/FAPAR model and the fesc_GPR-SR model with that simulated by the SCOPE model is shown in Figure 5, and it is evident that the fesc values estimated by the NIRv/FAPAR model are significantly underestimated (data points are below the 1:1 line), especially for canopies with low LAI values. Figure 6 clearly illustrates that as the LAI value increases, the underestimation effect of fesc calculated by the NIRv/FAPAR model decreases. The fesc estimated by the fesc_GPR-SR model is in close agreement with the true fesc values under all LAI values, which shows that the fesc_GPR-SR model can effectively correct the underestimation of fesc, especially for sparse vegetation. The R2 value of the relationship between fesc estimated by the fesc_GPR-SR model and the true fesc values improves from 0.730 to 0.949, and the RMSE reduces from 0.081 to 0.015, compared with the NIRv/FAPAR model. The improvement of the fesc_GPR-SR model is also significant in the case of sparse vegetation, with R2 increasing from 0.866 to 0.954 and RMSE decreasing from 0.100 to 0.012. The results demonstrate that the fesc_GPR-SR model provides a better estimation of fesc in the near-infrared band and has high estimation accuracy for sparse vegetation.
Furthermore, we calculated the coefficient of variation (CV) of the canopy SIF simulated by the SCOPE model and the leaf SIF estimated by the fesc_GPR-SR model and NIRv/FAPAR model for each set of different LAI values with only the VZA changing and the remaining parameters fixed (Figure 7). The smaller the CV value, the less the SIF is affected by the directional effect. The CV of canopy SIF for different LAI values is much larger than that of leaf SIF due to the anisotropic scattering within the canopy. Apparently, compared with the NIRv/FAPAR model, the CV of leaf SIF estimated by the fesc_GPR-SR model is smaller, and the gap between the results of the two models is more obvious for sparse vegetation. The result suggests that our fesc_GPR-SR model can reduce the directional effect of canopy SIF more effectively compared with the NIRv/FAPAR model.

3.2.2. Validation of the fesc_GPR-SR Model Using DART Simulations

The reliability of the fesc_GPR-SR model was further assessed using leaf-level and canopy-level SIF of maize plants simulated by the DART model under sparse vegetation conditions (LAI = 2). The comparison of SIF at canopy and leaf levels is shown in Figure 8. For ease of comparison, the unit of leaf-level SIF was converted from mW/m2/nm to mW/m2/nm/sr. The leaf-level SIF estimated by the fesc_GPR-SR model (the mean value is 1.818 mW/m2/nm/sr) is found to be closer to the SIF simulated by the DART model ( S I F l e a f D A R T = 1.813 mW/m2/nm/sr). The interquartile range (IQR) of leaf-level SIF calculated by the fesc_GPR-SR model is significantly smaller than that of multi-angle canopy SIF, indicating that the fesc_GPR-SR model can eliminate the directional effect of the canopy SIF effectively. Figure 9 displays the comparison of fesc in the near-infrared band estimated by the NIRv/FAPAR model and the fesc_GPR-SR model with that simulated by the DART model. The fesc estimated by the fesc_GPR-SR model is closer to the 1:1 line, with similar R2 values but the RMSE decreasing significantly from 0.070 to 0.026, in comparison with the NIRv/FAPAR model. These results of the 3-D radiation transfer model also verify that the fesc_GPR-SR model can effectively improve the accuracy of fesc estimation.

3.3. Evaluation of the fesc_GPR-SR Model Using Field-Measured Data

The ΦSIF in Equation (1) remains relatively unchanged under high-light and stress-free conditions [49,50]. As a result, there is a strong correlation between APARgreen and SIFtotal, which is stronger than at the canopy level, taking into account canopy scattering and reabsorption [15,26]. We use the field-measured data from healthy and non-stressed vegetation to study the correlation of APARgreen–SIF at both canopy and leaf levels to assess the performance of both downscaling methods.
To verify the improved performance of the model that accounts for soil reflectance in SIF downscaling, the samples with FVC ≤ 0.8 from the ground-measured dataset were selected for comparison and validation. Figure 10 reveals the correlation between APARgreen and the canopy-level SIF, the leaf-level SIF calculated by the NIRv/FAPAR model, and the fesc_GPR-SR model for various species. In comparison with the APARgreen–SIFcanopy relationship, the R2 of the correlation between SIF after downscaling using both models and APARgreen increases significantly and the slopes of the linear regression lines of APARgreen–SIFleaf for various species become more similar, indicating that downscaling SIF from canopy level to leaf level eliminates the impact of varying canopy structures among species and reduces the species dependence of the APARgreen–SIF correlation. Our fesc_GPR-SR model performs better in this aspect, as reflected by the fact that the R2 of the relationship between APARgreen and SIFleaf estimated by the fesc_GPR-SR model improves from 0.921 to 0.937 and the RMSE declines from 0.904 to 0.656 mW/m2/nm in comparison with the results of the NIRv/FAPAR model.
The R2 and RMSE values of the linear correlation between APARgreen and SIF for samples of various species with FVC ≤ 0.8 appear in Table 6. The results indicate that the R2 values of the APARgreen–SIFleaf relationship for different species increase after the SIF downscaling. Compared with the NIRv/FAPAR model, the RMSE of SIFleaf for different species calculated by our fesc_GPR-SR model is smaller, indicating that the fesc_GPR-SR model can better estimate the leaf-level SIF. For vegetables and crops, although the R2 value slightly decreases after accounting for soil reflectance, the RMSE is still smaller than that of the NIRv/FAPAR model. For gold coin grass and winter wheat, the SIFleaf calculated by the fesc_GPR-SR model displays a better linear correlation with APARgreen.
Overall, the outcomes indicate that the fesc_GPR-SR model shows better performance in SIF downscaling for different species, reducing the species dependence of APARgreen–SIF for sparse vegetation to a greater extent in contrast to the NIRv/FAPAR model.

4. Discussion

4.1. Effect of Soil Reflectance on Estimating fesc

The scattering and reabsorption of SIF photons within the canopy are governed by the same physical mechanisms as the scattering and reabsorption of reflected radiation photons. When incident light enters the canopy from the top, it can either pass through the canopy and reach the soil surface through the gaps or interact with leaves in the canopy. Photons intercepted by the canopy are scattered and reabsorbed several times, and some escape from the canopy while others are absorbed by leaves. Photons absorbed by leaves in the PAR range (400–700 nm) can stimulate fluorescence photons with a 640–850 nm wavelength range [27]. Emitted fluorescence will also escape from the canopy after multiple scatterings and reabsorptions inside the canopy and leaves.
In practice, soil background has a specific reflectance spectrum and is not “black.” Ignoring the influence of atmospheric radiative transfer and multiple scatterings between soil and canopy (considering only single scattering), the photons captured by the sensor mainly come from three sources: (1) photons (including emitted fluorescence photons) that escape upward from the canopy to the sensor; (2) incident photons that pass through the canopy, reach the soil surface, and are reflected to the sensor; (3) photons (including emitted fluorescence photons) escaping downward from the canopy that reach the soil and are reflected to the sensor. It is worth mentioning that the downward-escaping SIF photons from the canopy may be absorbed by the soil; however, this can be ignored as the lower leaves receive less light and thus produce fewer SIF photons [15]. Consequently, the SIF photons captured by the sensor come from two main sources: the contribution of pure vegetation canopy and the contribution of soil single scattering, and soil reflectance affects both. Yang et al. [27] proposed a basic model to estimate fesc based on spectrally invariant theory, where fesc = RefNIR/i0·ωN, ignoring the contribution of soil scattering to the fluorescence signal captured by the sensor. The near-infrared reflectance in the numerator clearly contains the effect of soil reflectance. Zeng et al. [28] proposed the NIRv/FAPAR model, which replaced the near-infrared reflectance with NIRv; however, the NDVI used as the pure vegetation signal in this model still depended on soil reflectance. The accuracy of fesc estimated by the NIRv/FAPAR model is reduced when a real soil background, rather than a non-reflecting background, is present, especially for sparse scenes [51]. This is because soil background pollutes the NIRv used to calculate fesc. The variability of the value of i0 is also high for sparse vegetation, which indicates that soil background can affect i0 and thus fesc [15].
The simulations using the SCOPE model in this study also show that soil reflectance significantly affects the estimation of fesc, particularly for sparse vegetation (Figure 11). When NIRv < 0.439, soil reflectance is the dominant factor in fesc, influencing the scattering process between the canopy and soil background. The canopy structure is the dominant factor of fesc when NIRv > 0.439. Thus, ignoring the reflection characteristic of the soil background and treating it as “black” will introduce significant uncertainty in the calculation of fesc, especially for sparse canopies where soil reflectance has a greater influence.

4.2. Superiority of the fesc_GPR-SR Model

At present, some studies on SIF downscaling have considered soil reflectance; however, these studies lacked clarity and had significant uncertainties. Liu et al. [15] proposed a SIF-downscaling approach where the ratio of fesc to the bi-directional reflectance factor was obtained through SCOPE model simulations and the hypothesis of “black soil” was ignored in the machine learning process. However, they did not explicitly account for the effect of soil reflectance. Zhang et al. [52] developed a method to derive the global soil-resistant SIFtotal (SIFtotal-SR) using satellite data. However, their method used an approximation of the observed minimum reflectance to soil reflectance that is uncertain and relied on LAI and clumping index (CI) satellite data, which also propagated uncertainties in SIFtotal. Notably, Zeng et al. [53] proposed using NIRvH with minimal soil impacts by making use of the spectral shape changes in the red-edge region to calculate true vegetation near-infrared reflectance. We used 9000 validation samples simulated by the SCOPE model to verify the performance of the NIRvH2/FAPAR model to estimate fesc (Figure S4). Compared with the results in Figure 5, it can be found that our model is also superior to the NIRvH2/FAPAR model (R2 = 0.885, RMSE = 0.051), possibly because of uncertainties in NIRvH itself, including the assumption that NIRvH2 is based on a linear increase in the reflectance of soil in the red-edge region which is not always correct.
Ma et al. [54] suggested that supervised machine learning methods trained on appropriate training datasets could construct accurate predictive models and overcome difficulties in physical modeling. Rasmussen et al. [55] also noted that while traditional parametric models are easy to interpret, they can be limited in their expression for complex datasets. Hence, in this study, we chose to build our fesc_GPR-SR model using a machine learning method. We proposed to add a correction factor composed of soil reflectance and NDVI to the widely used NIRv/FAPAR model suggested by Zeng et al. [28], in order to explicitly account for the effect of soil reflectance. This not only preserved the advantages of the NIRv/FAPAR model’s simplicity, ease of computation, and clear physical meaning but also compensated for its lack of explicit consideration of soil reflectance’s effect on fesc estimation. We tested the performance of various machine learning algorithms and found that exponential GPR was the best training model. Directly using the machine learning model to estimate fesc may lead to poor model robustness due to too many model input parameters, so the correction factor f (Refsoil, VI) with fewer parameters was only trained using the exponential GPR method, which helped to reduce the number of input parameters and improve the model’s robustness. Moreover, it ensured that the improvement was carried out on the basis of retaining the physical meaning of the original NIRv/FAPAR model as much as possible.
We demonstrated the superiority of our model by combining simulation data with field data. The validation results of SCOPE model simulations (Section 3.2.1) showed that the fesc estimated by the fesc_GPR-SR model was in good agreement with the true fesc values, even under sparse vegetation, and it can significantly eliminate the influence of direction effect. In addition, the 3-D radiative transfer model dataset also validated that the fesc_GPR-SR model could improve the estimation of fesc (Section 3.2.2). Since the DART model is too time-consuming to use for simulating a large amount of training data, a small dataset was used to evaluate the performance of the fesc estimation model trained with the SCOPE simulations. Additionally, ground-measured data from healthy and non-stressed vegetation were used for supplementary verification and showed that the linear correlation between APARgreen and SIFleaf could be improved for different species when soil reflectance is considered and FVC is less than or equal to 0.8. The fesc_GPR-SR model reduced the species dependency of the SIF–APARgreen relationship for sparse vegetation to a greater extent (Section 3.3). The R2 value of the APARgreen–SIFleaf correlation for vegetables and crops decreased slightly after considering soil reflectance, which may be because most of the photographs of vegetables and crops at Nanbin Farm were taken on cloudy days, leading to greater uncertainty in the FAPARgreen and soil reflectance calculations using the photographs. It must be noted that the significant linear relationship between APARgreen and SIF exists only in the absence of environmental stress. When environmental stress exists, APARgreen may not be able to characterize the leaf-level SIF. So, APARgreen cannot simply be used as a proxy of the total SIF. In conclusion, the fesc_GPR-SR model, which accounts for soil reflectance, has been verified to increase the accuracy of fesc estimation in the near-infrared band, particularly for sparse vegetation and the leaf-level SIF in the near-infrared band calculated by the fesc_GPR-SR model is less sensitive to observation angles and variations in canopy structure among multiple species.

4.3. Uncertainties of the fesc_GPR-SR Model

Although the fesc_GPR-SR model has been shown to be superior, there are still uncertainties in the modeling process. Firstly, in addition to the influence of soil background reflectance, the assumption that ωN was 1 also contributed to the underestimation of fesc by the NIRv/FAPAR model, because ωN is actually less than 1. However, considering that ωN is the sum of leaf reflectance and transmittance in the near-infrared band and the absorption effect of leaves in this band is very weak, numerous studies have shown and acknowledged that ωN is relatively stable and close to 1 [15,27,28,29]. Therefore, we did not pay attention to the uncertainty caused by the assumption that ωN was 1 but focused on the effect of soil reflectance on fesc estimation. In the future, improving the precision of fesc can be considered by calculating ωN accurately. Secondly, in this study, the correction factor was modeled using a combination of soil reflectance and the vegetation index NDVI, which represents vegetation coverage. While this combination was selected after evaluating the models’ performance using four common vegetation indices, the correction factor could be further optimized in the future by testing more than two multi-factor combinations. Thirdly, despite the computational advantages of machine learning methods, they are still black-box models and heavily dependent on training datasets, which reduces their adaptability in special conditions. In the future, we plan to improve the model by establishing a new semi-empirical analytical model that considers the influence of soil reflectance on fesc estimation in two parts (one is the influence on the calculation of pure vegetation reflectance and the other is the influence of single scattering between soil and canopy), based on the basic model (fesc = RefNIR/(i0·ωN)) proposed by Yang et al. [27]. Additionally, there are still uncertainties that cannot be clearly analyzed and quantified, including the retrieval error of SIF. In a word, there are still many problems to be addressed in future SIF-downscaling research.

5. Conclusions

Accurate estimation of the canopy fluorescence escaping probability is important for SIF application, but the current algorithms cannot well deal with the influence of soil reflectance, especially for sparse vegetation. In this work, a correction factor estimated using the GPR algorithm with soil reflectance and NDVI was introduced into the widely used NIRv/FAPAR model for better estimation of fesc in the near-infrared band. The new method we proposed, the fesc_GPR-SR model, was evaluated using simulation data and ground-measured data. The validation results of two simulation datasets from the SCOPE model and the DART model demonstrate that the performance of the fesc_GPR-SR model in estimating fesc is significantly better than that of the NIRv/FAPAR model, particularly for sparse vegetation, and our fesc_GPR-SR model can also effectively eliminate the influence of direction effects. Moreover, the validation results using the in situ measured data also prove that, compared with the NIRv/FAPAR model, our fesc_GPR-SR model can better reduce the species dependence of APARgreen–SIF and eliminate the effect of canopy structure difference in multiple species. This study highlights the significance and advantages of considering soil reflectance in fesc modeling and presents a more accurate fesc estimation model, which will be useful for further studies on the SIF–GPP relationship.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/rs15184361/s1, Figure S1: Soil spectra called “loam_gravelly_brown_dark” used in the DART model (from the DART model database “Lambertian.db”); Figure S2: (a) Nadir and (b, c) side view of the 3-D canopy of maize simulated using the DART model. The size of the scene is 1.5 m × 1 m, and 20 maize are planted in two rows, with 10 in each row. Figure S3: Training performance (Predicted vs. Actual) of machine learning algorithms based on 5-fold cross-validation method when the input parameters are soil reflectance at 780 nm and NDVI; Figure S4: Comparison of fesc in the near-infrared band (760 nm) estimated by the NIRvH2/FAPAR model with the fesc simulated by the SCOPE model. R2 is the correlation coefficient of linear regression, and RMSE is the root mean square error between fesc (SCOPE) and the fesc values calculated by the NIRvH2/FAPAR model; Table S1: The main parameters of the GPR model when using MATLAB to fit Gaussian process regression.

Author Contributions

Conceptualization, X.L. and L.L.; methodology, M.Q. and X.L.; software, M.Q.; formal analysis, M.Q.; investigation, M.Q.; resources, X.L., S.D. and L.L.; data curation, M.Q. and S.D.; writing—original draft preparation, M.Q.; writing—review and editing, X.L., L.L., L.G. and R.C.; visualization, M.Q.; funding acquisition, X.L. and L.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China, grant number 2022YFF1301900, and the National Natural Science Foundation of China, grant number 42071310.

Data Availability Statement

Data will be made available upon request.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Magney, T.S.; Bowling, D.R.; Logan, B.A.; Grossmann, K.; Stutz, J.; Blanken, P.D.; Burns, S.P.; Cheng, R.; Garcia, M.A.; Köhler, P.; et al. Mechanistic evidence for tracking the seasonality of photosynthesis with solar-induced fluorescence. Proc. Natl. Acad. Sci. USA 2019, 116, 11640–11645. [Google Scholar] [CrossRef] [PubMed]
  2. Porcar-Castell, A.; Tyystjärvi, E.; Atherton, J.; Van der Tol, C.; Flexas, J.; Pfündel, E.E.; Moreno, J.; Frankenberg, C.; Berry, J.A. Linking chlorophyll a fluorescence to photosynthesis for remote sensing applications: Mechanisms and challenges. J. Exp. Bot. 2014, 65, 4065–4095. [Google Scholar] [CrossRef] [PubMed]
  3. He, L.; Magney, T.; Dutta, D.; Yin, Y.; Köhler, P.; Grossmann, K.; Stutz, J.; Dold, C.; Hatfield, J.; Guan, K. From the ground to space: Using solar-induced chlorophyll fluorescence to estimate crop productivity. Geophys. Res. Lett. 2020, 47, e2020GL087474. [Google Scholar] [CrossRef]
  4. Kimm, H.; Guan, K.; Burroughs, C.H.; Peng, B.; Ainsworth, E.A.; Bernacchi, C.J.; Moore, C.E.; Kumagai, E.; Yang, X.; Berry, J.A.; et al. Quantifying high-temperature stress on soybean canopy photosynthesis: The unique role of sun-induced chlorophyll fluorescence. Glob. Chang. Biol. 2021, 27, 2403–2415. [Google Scholar] [CrossRef]
  5. Migliavacca, M.; Perez-Priego, O.; Rossini, M.; El-Madany, T.S.; Moreno, G.; Van der Tol, C.; Rascher, U.; Berninger, A.; Bessenbacher, V.; Burkart, A. Plant functional traits and canopy structure control the relationship between photosynthetic CO2 uptake and far-red sun-induced fluorescence in a Mediterranean grassland under different nutrient availability. New Phytol. 2017, 214, 1078–1091. [Google Scholar] [CrossRef]
  6. Sun, Y.; Frankenberg, C.; Wood, J.D.; Schimel, D.S.; Jung, M.; Guanter, L.; Drewry, D.T.; Verma, M.; Porcar-Castell, A.; Griffis, T.J.; et al. OCO-2 advances photosynthesis observation from space via solar-induced chlorophyll fluorescence. Science 2017, 358, eaam5747. [Google Scholar] [CrossRef]
  7. Liu, Z.; Zhao, F.; Liu, X.; Yu, Q.; Wang, Y.; Peng, X.; Cai, H.; Lu, X. Direct estimation of photosynthetic CO2 assimilation from solar-induced chlorophyll fluorescence (SIF). Remote Sens. Environ. 2022, 271, 112893. [Google Scholar] [CrossRef]
  8. Mohammed, G.H.; Colombo, R.; Middleton, E.M.; Rascher, U.; Zarco-Tejada, P.J. Remote sensing of solar-induced chlorophyll fluorescence (SIF) in vegetation: 50 years of progress. Remote Sens. Environ. 2019, 231, 111177. [Google Scholar] [CrossRef]
  9. Porcar-Castell, A.; Malenovský, Z.; Magney, T.S.; Van Wittenberghe, S.; Fernández-Marín, B.; Maignan, F.; Zhang, Y.; Maseyk, K.; Atherton, J.; Albert, L.P.; et al. Chlorophyll a fluorescence illuminates a path connecting plant molecular biology to Earth-system science. Nat. Plants 2021, 7, 998–1009. [Google Scholar] [CrossRef]
  10. Frankenberg, C.; O’Dell, C.; Berry, J.; Guanter, L.; Joiner, J.; Köhler, P.; Pollock, R.; Taylor, T.E. Prospects for chlorophyll fluorescence remote sensing from the Orbiting Carbon Observatory-2. Remote Sens. Environ. 2014, 147, 1–12. [Google Scholar] [CrossRef]
  11. Gu, L.; Han, J.; Wood, J.D.; Chang, C.Y.Y.; Sun, Y. Sun-induced Chl fluorescence and its importance for biophysical modeling of photosynthesis based on light reactions. New Phytol. 2019, 223, 1179–1191. [Google Scholar] [CrossRef] [PubMed]
  12. Du, S.; Liu, X.; Chen, J.; Liu, L. Prospects for Solar-Induced Chlorophyll Fluorescence Remote Sensing from the SIFIS Payload Onboard the TECIS-1 Satellite. J. Remote Sens. 2022, 2022, 9845432. [Google Scholar] [CrossRef]
  13. Frankenberg, C.; Berry, J.; Guanter, L.; Joiner, J. Remote sensing of terrestrial chlorophyll fluorescence from space. SPIENewsroom 2013, 2–5, 4725. [Google Scholar] [CrossRef]
  14. Frankenberg, C.; Fisher, J.B.; Worden, J.; Badgley, G.; Saatchi, S.S.; Lee, J.-E.; Toon, G.C.; Butz, A.; Jung, M.; Kuze, A.; et al. New global observations of the terrestrial carbon cycle from GOSAT: Patterns of plant fluorescence with gross primary productivity. Geophys. Res. Lett. 2011, 38, L17706. [Google Scholar] [CrossRef]
  15. Liu, X.; Guanter, L.; Liu, L.; Damm, A.; Malenovský, Z.; Rascher, U.; Peng, D.; Du, S.; Gastellu-Etchegorry, J.-P. Downscaling of solar-induced chlorophyll fluorescence from canopy level to photosystem level using a random forest model. Remote Sens. Environ. 2019, 231, 110772. [Google Scholar] [CrossRef]
  16. Li, Z.; Zhang, Q.; Li, J.; Yang, X.; Wu, Y.; Zhang, Z.; Wang, S.; Wang, H.; Zhang, Y. Solar-induced chlorophyll fluorescence and its link to canopy photosynthesis in maize from continuous ground measurements. Remote Sens. Environ. 2020, 236, 111420. [Google Scholar] [CrossRef]
  17. Hao, D.; Zeng, Y.; Qiu, H.; Biriukova, K.; Celesti, M.; Migliavacca, M.; Rossini, M.; Asrar, G.R.; Chen, M. Practical approaches for normalizing directional solar-induced fluorescence to a standard viewing geometry. Remote Sens. Environ. 2021, 255, 112171. [Google Scholar] [CrossRef]
  18. Siegmann, B.; Cendrero-Mateo, M.P.; Cogliati, S.; Damm, A.; Gamon, J.; Herrera, D.; Jedmowski, C.; Junker-Frohn, L.V.; Kraska, T.; Muller, O. Downscaling of far-red solar-induced chlorophyll fluorescence of different crops from canopy to leaf level using a diurnal data set acquired by the airborne imaging spectrometer HyPlant. Remote Sens. Environ. 2021, 264, 112609. [Google Scholar] [CrossRef]
  19. van der Tol, C.; Rossini, M.; Cogliati, S.; Verhoef, W.; Colombo, R.; Rascher, U.; Mohammed, G. A model and measurement comparison of diurnal cycles of sun-induced chlorophyll fluorescence of crops. Remote Sens. Environ. 2016, 186, 663–677. [Google Scholar] [CrossRef]
  20. Berry, J.A.; Frankenberg, C.; Wennberg, P.; Baker, I.; Bowman, K.W.; Castro-Contreas, S.; Cendrero-Mateo, M.P.; Damm, A.; Drewry, D.; Ehlmann, B. New Methods for Measurement of Photosynthesis from Space. Geophys. Res. Lett. 2012, 38, L17706. [Google Scholar]
  21. Liu, L.; Guan, L.; Liu, X. Directly estimating diurnal changes in GPP for C3 and C4 crops using far-red sun-induced chlorophyll fluorescence. Agric. For. Meteorol. 2017, 232, 1–9. [Google Scholar] [CrossRef]
  22. Liu, L.; Liu, X.; Hu, J.; Guan, L. Assessing the wavelength-dependent ability of solar-induced chlorophyll fluorescence to estimate the GPP of winter wheat at the canopy level. Int. J. Remote Sens. 2017, 38, 4396–4417. [Google Scholar] [CrossRef]
  23. Wieneke, S.; Ahrends, H.; Damm, A.; Pinto, F.; Stadler, A.; Rossini, M.; Rascher, U. Airborne based spectroscopy of red and far-red sun-induced chlorophyll fluorescence: Implications for improved estimates of gross primary productivity. Remote Sens. Environ. 2016, 184, 654–667. [Google Scholar] [CrossRef]
  24. Zhang, Y.; Xiao, X.; Wolf, S.; Wu, J.; Wu, X.; Gioli, B.; Wohlfahrt, G.; Cescatti, A.; Van der Tol, C.; Zhou, S. 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]
  25. Zhang, Z.; Zhang, Y.; Zhang, Y.; Gobron, N.; Frankenberg, C.; Wang, S.; Li, Z. The potential of satellite FPAR product for GPP estimation: An indirect evaluation using solar-induced chlorophyll fluorescence. Remote Sens. Environ. 2020, 240, 111686. [Google Scholar] [CrossRef]
  26. Liu, X.; Liu, L.; Hu, J.; Guo, J.; Du, S. Improving the potential of red SIF for estimating GPP by downscaling from the canopy level to the photosystem level. Agric. For. Meteorol. 2020, 281, 107846. [Google Scholar] [CrossRef]
  27. Yang, P.; van der Tol, C. Linking canopy scattering of far-red sun-induced chlorophyll fluorescence with reflectance. Remote Sens. Environ. Interdiscip. J. 2018, 209, 456–467. [Google Scholar] [CrossRef]
  28. Zeng, Y.; Badgley, G.; Dechant, B.; Ryu, Y.; Berry, J.A. A practical approach for estimating the escape ratio of near-infrared solar-induced chlorophyll fluorescence. Remote Sens. Environ. 2019, 232, 111209. [Google Scholar] [CrossRef]
  29. Yang, P.; van der Tol, C.; Campbell, P.K.E.; Middleton, E.M. Fluorescence Correction Vegetation Index (FCVI): A physically based reflectance index to separate physiological and non-physiological information in far-red sun-induced chlorophyll fluorescence. Remote Sens. Environ. 2020, 240, 111676. [Google Scholar] [CrossRef]
  30. van der Tol, C.; Verhoef, W.; Timmermans, J.; Verhoef, A.; Su, Z. An integrated model of soil-canopy spectral radiances, photosynthesis, fluorescence, temperature and energy balance. Biogeosciences 2009, 6, 3109–3129. [Google Scholar] [CrossRef]
  31. Jacquemoud, S.; Baret, F. PROSPECT: A model of leaf optical properties spectra. Remote Sens. Environ. 1990, 34, 75–91. [Google Scholar] [CrossRef]
  32. Verhoef, W. Light scattering by leaf layers with application to canopy reflectance modeling: The SAIL model. Remote Sens. Environ. 1984, 16, 125–141. [Google Scholar] [CrossRef]
  33. Qian, X.; Liu, L.; Chen, X.; Zhang, X.; Chen, S.; Sun, Q. The global leaf chlorophyll content dataset over 2003–2012 and 2018–2020 derived from MERIS/OLCI satellite data (GLCC): Algorithm and validation. Earth Syst. Sci. Data Discuss. 2022, 2022, 1–21. [Google Scholar] [CrossRef]
  34. Gastellu-Etchegorry, J.P.; Martin, E.; Gascon, F. DART: A 3D model for simulating satellite images and studying surface radiation budget. Int. J. Remote Sens. 2004, 25, 73–96. [Google Scholar] [CrossRef]
  35. Gastellu-Etchegorry, J.-P.; Lauret, N.; Yin, T.; Landier, L.; Kallel, A.; Malenovský, Z.; Al Bitar, A.; Aval, J.; Hmida, S.B.; Qi, J.; et al. DART: Recent Advances in Remote Sensing Data Modeling With Atmosphere, Polarization, and Chlorophyll Fluorescence. IEEE J. Sel. Top. Appl. Earth Obs. Remote Sens. 2017, 10, 2640–2649. [Google Scholar] [CrossRef]
  36. Maier, S.W.; Günther, K.; Stellmes, M. Sun-Induced Fluorescence: A New Tool for Precision Farming; ASA: Madison, WI, USA; Monroe, MI, USA, 2003. [Google Scholar]
  37. Damm, A.; Erler, A.; Hillen, W.; Meroni, M.; Schaepman, M.E.; Verhoef, W.; Rascher, U. Modeling the impact of spectral sensor configurations on the FLD retrieval accuracy of sun-induced chlorophyll fluorescence. Remote Sens. Environ. 2011, 115, 1882–1892. [Google Scholar] [CrossRef]
  38. Liu, L.; Liu, X.; Hu, J. Effects of spectral resolution and SNR on the vegetation solar-induced fluorescence retrieval using FLD-based methods at canopy level. Eur. J. Remote Sens. 2015, 48, 743–762. [Google Scholar] [CrossRef]
  39. Gitelson, A.A. Remote estimation of fraction of radiation absorbed by photosynthetically active vegetation: Generic algorithm for maize and soybean. Remote Sens. Lett. 2018, 10, 283–291. [Google Scholar] [CrossRef]
  40. Liu, L.; Peng, D.; Hu, Y.; Jiao, Q. A novel in situ FPAR measurement method for low canopy vegetation based on a digital camera and reference panel. Remote Sens. 2013, 5, 274–281. [Google Scholar] [CrossRef]
  41. Rouse, J.W.; Haas, R.H.; Schell, J.A.; Deering, D.W. Monitoring vegetation systems in the Great Plains with ERTS. NASA Spec. Publ. 1974, 351, 309. [Google Scholar]
  42. Jordan, C.F. Derivation of Leaf-Area Index from Quality of Light on the Forest Floor. Ecology 1969, 50, 663–666. [Google Scholar] [CrossRef]
  43. Jiang, Z.; Huete, A.R.; Didan, K.; Miura, T. Development of a two-band enhanced vegetation index without a blue band. Remote Sens. Environ. 2008, 112, 3833–3845. [Google Scholar] [CrossRef]
  44. Richardsons, A.J.; Wiegand, A. Distinguishing vegetation from soil background information. Photogramm. Eng. Remote Sens. 1977, 43, 1541–1552. [Google Scholar] [CrossRef]
  45. Schulz, E.; Speekenbrink, M.; Krause, A. A tutorial on Gaussian process regression: Modelling, exploring, and exploiting functions. J. Math. Psychol. 2018, 85, 1–16. [Google Scholar] [CrossRef]
  46. Jin, H.; Chen, X.; Wang, Y.; Zhong, R.; Zhao, T.; Liu, Z.; Tu, X. Spatio-temporal distribution of NDVI and its influencing factors in China. J. Hydrol. 2021, 603, 127129. [Google Scholar] [CrossRef]
  47. Jiang, Z.; Huete, A.R.; Chen, J.; Chen, Y.; Li, J.; Yan, G.; Zhang, X. Analysis of NDVI and scaled difference vegetation index retrievals of vegetation fraction. Remote Sens. Environ. 2006, 101, 366–378. [Google Scholar] [CrossRef]
  48. Zeng, Y.; Hao, D.; Huete, A.; Dechant, B.; Berry, J.; Chen, J.M.; Joiner, J.; Frankenberg, C.; Bond-Lamberty, B.; Ryu, Y. Optical vegetation indices for monitoring terrestrial ecosystems globally. Nat. Rev. Earth Environ. 2022, 3, 477–493. [Google Scholar] [CrossRef]
  49. Damm, A.; Guanter, L.; Paul-Limoges, E.; van der Tol, C.; Hueni, A.; Buchmann, N.; Eugster, W.; Ammann, C.; Schaepman, M.E. Far-red sun-induced chlorophyll fluorescence shows ecosystem-specific relationships to gross primary production: An assessment based on observational and modeling approaches. Remote Sens. Environ. 2015, 166, 91–105. [Google Scholar] [CrossRef]
  50. van der Tol, C.; Berry, J.A.; Campbell, P.K.E.; Rascher, U. Models of fluorescence and photosynthesis for interpreting measurements of solar-induced chlorophyll fluorescence. J. Geophys. Res. Biogeosci. 2014, 119, 2312–2327. [Google Scholar] [CrossRef]
  51. Liu, W.; Luo, S.; Lu, X.; Atherton, J.; Gastellu-Etchegorry, J.-P. Simulation-based evaluation of the estimation methods of far-red solar-induced chlorophyll fluorescence escape probability in discontinuous Forest canopies. Remote Sens. 2020, 12, 3962. [Google Scholar] [CrossRef]
  52. Zhang, Z.; Chen, J.M.; Guanter, L.; He, L.; Zhang, Y. From canopy-leaving to total canopy far-red fluorescence emission for remote sensing of photosynthesis: First results from TROPOMI. Geophys. Res. Lett. 2019, 46, 12030–12040. [Google Scholar] [CrossRef]
  53. Zeng, Y.; Hao, D.; Badgley, G.; Damm, A.; Rascher, U.; Ryu, Y.; Johnson, J.; Krieger, V.; Wu, S.; Qiu, H.; et al. Estimating near-infrared reflectance of vegetation from hyperspectral data. Remote Sens. Environ. 2021, 267, 112723. [Google Scholar] [CrossRef]
  54. Ma, C.; Zhang, H.H.; Wang, X. Machine learning for Big Data analytics in plants. Trends Plant Sci. 2014, 19, 798–808. [Google Scholar] [CrossRef] [PubMed]
  55. Rasmussen, C.E.; Williams, C.K. Gaussian processes in machine learning. Lect. Notes Comput. Sci. 2004, 3176, 63–71. [Google Scholar]
Figure 1. Soil spectral curves used in the SCOPE model. The black dashed line shows the wavelength position at 780 nm.
Figure 1. Soil spectral curves used in the SCOPE model. The black dashed line shows the wavelength position at 780 nm.
Remotesensing 15 04361 g001
Figure 2. Multi-angle canopy-level SIF results of maize canopies in the near-infrared band (760 nm) using the DART model. The angular coordinate and diameter coordinate represent the viewing azimuth angles (VAAs, 0–360°) and the viewing zenith angles (VZAs, 0–90°), respectively. The red pentagram indicates the position of the sun (SZA = 30.9303°, SAA = 249.1069°).
Figure 2. Multi-angle canopy-level SIF results of maize canopies in the near-infrared band (760 nm) using the DART model. The angular coordinate and diameter coordinate represent the viewing azimuth angles (VAAs, 0–360°) and the viewing zenith angles (VZAs, 0–90°), respectively. The red pentagram indicates the position of the sun (SZA = 30.9303°, SAA = 249.1069°).
Remotesensing 15 04361 g002
Figure 3. Photographs of canopies at (a) XTS in December (winter wheat), (b) NBF (sweet potato), and (c) SYS (gold coin grass). Only one representative photo is shown at each site here.
Figure 3. Photographs of canopies at (a) XTS in December (winter wheat), (b) NBF (sweet potato), and (c) SYS (gold coin grass). Only one representative photo is shown at each site here.
Remotesensing 15 04361 g003
Figure 4. Comparison of the determination coefficients R2 of exponential GPR models with different input and output parameter combinations. The horizontal coordinate represents the inputs of the model, and the color of the bars represents the output of the model. Light blue and light orange bars indicate the results with LAI < 2.
Figure 4. Comparison of the determination coefficients R2 of exponential GPR models with different input and output parameter combinations. The horizontal coordinate represents the inputs of the model, and the color of the bars represents the output of the model. Light blue and light orange bars indicate the results with LAI < 2.
Remotesensing 15 04361 g004
Figure 5. Comparison of fesc in the near-infrared band (760 nm) estimated by (a) the NIRv/FAPAR model (fesc (NIRv/FAPAR)) and (b) the fesc_GPR-SR model (fesc (GPR-SR)) with the fesc simulated by the SCOPE model. R2 is the correlation coefficient of linear regression, and RMSE is the root mean square error between fesc (SCOPE) and the fesc values calculated by fesc estimation models.
Figure 5. Comparison of fesc in the near-infrared band (760 nm) estimated by (a) the NIRv/FAPAR model (fesc (NIRv/FAPAR)) and (b) the fesc_GPR-SR model (fesc (GPR-SR)) with the fesc simulated by the SCOPE model. R2 is the correlation coefficient of linear regression, and RMSE is the root mean square error between fesc (SCOPE) and the fesc values calculated by fesc estimation models.
Remotesensing 15 04361 g005
Figure 6. Comparison of fesc in the near-infrared band (760 nm) calculated by the NIRv/FAPAR model (fesc (NIRv/FAPAR), orange boxes), the SCOPE model (fesc (SCOPE), green boxes), and the fesc_GPR-SR model (fesc (GPR-SR), blue boxes) under different LAI values.
Figure 6. Comparison of fesc in the near-infrared band (760 nm) calculated by the NIRv/FAPAR model (fesc (NIRv/FAPAR), orange boxes), the SCOPE model (fesc (SCOPE), green boxes), and the fesc_GPR-SR model (fesc (GPR-SR), blue boxes) under different LAI values.
Remotesensing 15 04361 g006
Figure 7. The coefficients of variation (CVs) of the canopy SIF simulated by the SCOPE model and the leaf SIF estimated by the fesc_GPR-SR model and NIRv/FAPAR model for each set of different LAI values with only the VZA changing and the remaining parameters fixed (LIDF is plagiophile, Cab = 60 μg/cm2, SZA = 30°, RAA = 180°, and soil reflectance at 780 nm is 0.1508). The amount of data for different LAI values is the same.
Figure 7. The coefficients of variation (CVs) of the canopy SIF simulated by the SCOPE model and the leaf SIF estimated by the fesc_GPR-SR model and NIRv/FAPAR model for each set of different LAI values with only the VZA changing and the remaining parameters fixed (LIDF is plagiophile, Cab = 60 μg/cm2, SZA = 30°, RAA = 180°, and soil reflectance at 780 nm is 0.1508). The amount of data for different LAI values is the same.
Remotesensing 15 04361 g007
Figure 8. Boxplot of maize canopy SIF in the near-infrared band (760 nm) simulated by the DART model (SIFcanopy) and leaf-level SIF estimated by the NIRv/FAPAR model ( S I F l e a f N I R v / F A P A R ) and that estimated by the fesc_GPR-SR model ( S I F l e a f G P R S R ). The red dotted lines represent the leaf-level SIF of maize in the near-infrared band (760 nm) simulated by the DART model ( S I F l e a f D A R T ). For the convenience of comparison, the unit of leaf-level SIF is converted to mW/m2/nm/sr (consistent with the unit of canopy SIF).
Figure 8. Boxplot of maize canopy SIF in the near-infrared band (760 nm) simulated by the DART model (SIFcanopy) and leaf-level SIF estimated by the NIRv/FAPAR model ( S I F l e a f N I R v / F A P A R ) and that estimated by the fesc_GPR-SR model ( S I F l e a f G P R S R ). The red dotted lines represent the leaf-level SIF of maize in the near-infrared band (760 nm) simulated by the DART model ( S I F l e a f D A R T ). For the convenience of comparison, the unit of leaf-level SIF is converted to mW/m2/nm/sr (consistent with the unit of canopy SIF).
Remotesensing 15 04361 g008
Figure 9. Comparison of fesc in the near-infrared band (760 nm) simulated by the DART model (fesc (DART)) with that estimated by the NIRv/FAPAR model (fesc (NIRv/FAPAR)) and the fesc_GPR-SR model (fesc (GPR-SR)). R2 is the correlation coefficient of linear regression, and RMSE is the root mean square error between fesc (DART) and the fesc values calculated by fesc estimation models.
Figure 9. Comparison of fesc in the near-infrared band (760 nm) simulated by the DART model (fesc (DART)) with that estimated by the NIRv/FAPAR model (fesc (NIRv/FAPAR)) and the fesc_GPR-SR model (fesc (GPR-SR)). R2 is the correlation coefficient of linear regression, and RMSE is the root mean square error between fesc (DART) and the fesc values calculated by fesc estimation models.
Remotesensing 15 04361 g009
Figure 10. Correlation between APARgreen and (a) canopy-level SIF (SIFcanopy), (b) leaf-level SIF estimated by the NIRv/FAPAR model ( S I F l e a f N I R v / F A P A R ), and (c) leaf-level SIF estimated by the fesc_GPR-SR model ( S I F l e a f G P R S R ) for different species (vegetables and crops, gold coin grass, and winter wheat) in the case of FVC ≤ 0.8. The black solid lines and the equations are the linear regression lines and models for all samples with FVC ≤ 0.8. The finer colored lines are the linear regression lines of the species that correspond in color.
Figure 10. Correlation between APARgreen and (a) canopy-level SIF (SIFcanopy), (b) leaf-level SIF estimated by the NIRv/FAPAR model ( S I F l e a f N I R v / F A P A R ), and (c) leaf-level SIF estimated by the fesc_GPR-SR model ( S I F l e a f G P R S R ) for different species (vegetables and crops, gold coin grass, and winter wheat) in the case of FVC ≤ 0.8. The black solid lines and the equations are the linear regression lines and models for all samples with FVC ≤ 0.8. The finer colored lines are the linear regression lines of the species that correspond in color.
Remotesensing 15 04361 g010
Figure 11. Relationship between fesc and NIRv simulated by the SCOPE model. The small circles in the figure represent the average values of different NIRv segments. The six curves in different colors represent the six soil spectral curves used in the SCOPE simulations.
Figure 11. Relationship between fesc and NIRv simulated by the SCOPE model. The small circles in the figure represent the average values of different NIRv segments. The six curves in different colors represent the six soil spectral curves used in the SCOPE simulations.
Remotesensing 15 04361 g011
Table 1. Main input variables of the SCOPE model.
Table 1. Main input variables of the SCOPE model.
VariablesDefinitionValuesUnit
CabLeaf chlorophyll a and b content20, 40, 60, 80μg/cm2
LAILeaf area index0.25, 0.5, 0.75, 1, 1.5, 2, 3, 5, 7m2/m2
LIDFaLeaf inclination parameter1, 0, 0, −0.35, 0
LIDFbBimodality parameter0, −1, 1, −0.15, 0
SZASolar zenith angle20, 30, 40, 50, 60Degree
VZAViewing zenith angle0, 15, 30, 45, 60Degree
RAARelative azimuth angle0, 90, 180Degree
Soil spectraSoil reflectanceSix soil spectra
Table 2. Main input variables of the DART model.
Table 2. Main input variables of the DART model.
VariablesDefinitionValuesUnit
Vegetation typeVegetation typeMaize
NStructure coefficient1.5
CabLeaf chlorophyll a and b content58μg/cm2
Yield PSIFluorescence quantum yield for photosystem I0.002
Yield PSIIFluorescence quantum yield for photosystem II0.008
LAILeaf area index2m2/m2
Canopy heightCanopy height1.5m
Soil spectraSoil reflectanceloam_gravelly_brown_dark
SZASolar zenith angle30.9303Degree
SAASolar azimuth angle249.1069Degree
VZAViewing zenith angle0–90Degree
VAAViewing azimuth angle0–360Degree
Table 3. Detailed information of the four ground measurements.
Table 3. Detailed information of the four ground measurements.
SitesXiaotangshan FarmXiaotangshan FarmNanbin FarmSanya Station
Location40°11′N
116°27′E
40°11′N
116°27′E
18°22′N
109°10′E
18°18′N
109°18′E
Dates in 20168, 9, 18 April8 December18 December18 December
SpeciesWinter wheatWinter wheatVegetables and cropsGold coin grass
Fractional vegetation cover (FVC)0.72–0.790.21–0.630.28–0.910.67
Soil reflectance in NIR band *0.12–0.170.13–0.160.09–0.160.11–0.13
* Soil reflectance in NIR band included that of litters.
Table 4. VIs used for the calculation of the correction factor (R780, R710, and R678 represent the reflectance at 780 nm, 710 nm, and 678 nm, respectively).
Table 4. VIs used for the calculation of the correction factor (R780, R710, and R678 represent the reflectance at 780 nm, 710 nm, and 678 nm, respectively).
VIsReferences
N D V I = ( R 780 R 678 ) / ( R 780 + R 678 ) [41]
S R = R 780 / R 678 [42]
E V I 2 = 2.5 × ( R 780 R 710 ) R 780 + 2.4 × R 678 + 1 [43]
P V I = ( R 780 s o i l R 780 v e g ) 2 + ( R 678 s o i l R 678 v e g ) 2 [44]
Table 5. Statistics of performance evaluation indices of the exponential GPR method with different input parameters.
Table 5. Statistics of performance evaluation indices of the exponential GPR method with different input parameters.
Inputs for ModelsR2RMSEMAER2 (LAI < 2)RMSE (LAI < 2)MAE (LAI < 2)
Refsoil, NDVI0.850.03910.03110.690.04380.0348
Refsoil, SR0.850.03960.03150.680.04410.0350
Refsoil, EVI20.840.04030.03130.650.04670.0372
Refsoil, PVI0.820.04330.03340.620.04840.0385
Table 6. R2 and RMSE values of the line correlation between APARgreen and SIF for samples of various species with FVC ≤ 0.8. The units of RMSE for the APARgreen–SIFcanopy relationship and the APARgreen–SIFleaf relationship are mW/m2/nm/sr and mW/m2/nm, respectively.
Table 6. R2 and RMSE values of the line correlation between APARgreen and SIF for samples of various species with FVC ≤ 0.8. The units of RMSE for the APARgreen–SIFcanopy relationship and the APARgreen–SIFleaf relationship are mW/m2/nm/sr and mW/m2/nm, respectively.
Vegetables and CropsGold Coin GrassWinter Wheat
R2RMSER2RMSER2RMSE
S I F c a n o p y 0.8830.187 0.7500.1950.9440.071
S I F l e a f N I R v / F A P A R 0.9550.7090.8910.6120.9660.683
S I F l e a f G P R S R 0.9500.6460.9010.5180.9700.511
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

Qi, M.; Liu, X.; Du, S.; Guan, L.; Chen, R.; Liu, L. Improving the Estimation of Canopy Fluorescence Escape Probability in the Near-Infrared Band by Accounting for Soil Reflectance. Remote Sens. 2023, 15, 4361. https://doi.org/10.3390/rs15184361

AMA Style

Qi M, Liu X, Du S, Guan L, Chen R, Liu L. Improving the Estimation of Canopy Fluorescence Escape Probability in the Near-Infrared Band by Accounting for Soil Reflectance. Remote Sensing. 2023; 15(18):4361. https://doi.org/10.3390/rs15184361

Chicago/Turabian Style

Qi, Mengjia, Xinjie Liu, Shanshan Du, Linlin Guan, Ruonan Chen, and Liangyun Liu. 2023. "Improving the Estimation of Canopy Fluorescence Escape Probability in the Near-Infrared Band by Accounting for Soil Reflectance" Remote Sensing 15, no. 18: 4361. https://doi.org/10.3390/rs15184361

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