Open Access
19 June 2017 Using remotely sensed spectral reflectance to indicate leaf photosynthetic efficiency derived from active fluorescence measurements
Yi Peng, Aoli Zeng, Tinge Zhu, Shenghui Fang, Yan Gong, Yanqi Tao, Ying Zhou, Kan Liu
Author Affiliations +
Abstract
Chlorophyll fluorescence (ChlF) is an important signature of photosynthesis to evaluate plant response to the environment. We explored an approach to estimate an important leaf ChlF-derived parameter, the intrinsic efficiency of photosystem II photochemistry ( F v / F m ), using spectral indices calculated from leaf reflectance measured by a hyperspectral radiometer. It is observed that leaf chlorophyll content closely related to F v / F m in nonstressed leaves, thus the indices developed for chlorophyll estimation were successfully used to estimate F v / F m . For leaves under short-term stress, F v / F m dropped dramatically while leaf chlorophyll content remained almost the same. Compared to leaf chlorophyll content, reflectance was more sensitive to F v / F m variations. As F v / F m decreased, the slope of reflectance in the spectrum range of 700 to 900 nm obviously increased, and the first derivative reflectance in the red edge and infrared (NIR) regions was highly correlated with

1.

Introduction

Chlorophyll fluorescence (ChlF) is the red and far-red light emitted by chlorophyll pigments upon excitation with natural or artificial light in the visible and infrared (NIR) part of the spectrum.1 It is widely realized that ChlF carries valuable information for the organization, functioning, and physiology of plant photosynthesis.2,3 Energy absorbed by chlorophyll of photosystem II (PSII) is used to split water into hydrogen and oxygen, but some can be dissipated as heat or re-emitted at longer wavelengths as ChlF. The capture of ChlF variations helps understand the processes that control energy partitioning in PSII, which is useful to evaluate plant growth and stress status.4 Many studies found that ChlF is a good indicator for terrestrial gross primary production, fraction of absorbed photosynthetically active radiation (PAR), light use efficiency, and maximum carboxylation capacity.57 Moreover, the first response of a plant to environmental stress is usually an increase of nonradiative energy dissipation likely indicated by ChlF.8 Thus, ChlF usually acts as a sensitive probe for early detection of plant stress.9

The pulse amplitude-modulated (PAM) technique is widely used for leaf ChlF measurements. It employs rapid manipulations of the light regime with modulated measuring light to obtain leaf fluorescence signals under different conditions.4 The measurement is initiated with a measuring light on a dark-adapted leaf, giving the minimal level of fluorescence (F0) when all photochemical traps open and nonphotochemical quenching (NPQ) approaches zero. A saturating pulse of light is then applied to close nearly all photochemical traps thus fluorescence rapidly increases to a maximal level (Fm). Turning on an actinic light can activate photosynthesis in leaf, and after a period of time, leaf fluorescence comes to a steady-state level (Ft). By applying saturating pulses, the maximal level of fluorescence in the light (Fm) can be measured. The removal of actinic light after a flash gives the minimal level of fluorescence (F0) in the light. Based on these recorded signals, various important parameters can be derived to quantify leaf photochemical and nonphotochemical processes, such as the efficiency of PSII photochemistry [ϕPSII=(FmFt)/Fm], photochemical quenching [qP=(FmFt)/(FmF0)], intrinsic efficiency of PSII photochemistry [Fv/Fm=(FmF0)/Fm], NPQ [NPQ=(FmFm)/Fm], and so on.4,10 The development of the PAM technique has greatly facilitated the study and applications of using ChlF information to evaluate vegetation photosynthesis machinery and physiology status. Several PAM-based fluorometers have been invented to obtain ChlF information in situ,11 but they are commonly applied at leaf level and their applicability at canopy or landscape level still remains elusive.4,12

Recently, estimating ChlF from remote sensing platforms is of great interest. Unlike PAM techniques based on fluorescence measurements induced by an active light source, remote sensing approaches retrieve ChlF excited by the absorption of sunlight [i.e., solar-induced fluorescence (SIF)] using spectral reflectance. Since the fluorescence emission spectrum is superimposed on leaf or canopy reflectance that can be obtained by handheld, ground-mounted, aerial-, or space-borne sensors, remote sensing technique opens the new way for upscaling ChlF from leaf to landscape levels. For example, Moya et al.13 designed a passive instrument to detect SIF at close range for leaves in lab and for outdoor plants; Rascher et al.14 mapped the relative distribution of SIF for maize fields using spectroscope aboard aircraft; and Joiner et al.15 achieved the first observation of global terrestrial SIF using the Japanese Greenhouse Gases Observing Satellite data. Also, the European Space Agency has started the Fluorescence Explorer mission specifically for ChlF assessments from space.16,17 The accurate quantification of regional and global ChlF has attracted wide attention, since it is a complementary and high capacity signal for evaluating terrestrial production and global carbon cycle.6,18

Since ChlF emitted by plant under natural sunlight is very small (only 1% to 2% of the absorbed light),19 it is quite challenging to extract such weak signal from much higher detected reflectance signal. Fraunhofer line discrimination (FLD) approach was developed to retrieve SIF at the distance from close range to satellite altitudes.20 Fraunhofer lines are narrow bands at which the incident solar irradiance is strongly reduced by gaseous absorption. Several Fraunhofer lines are observed at wavelengths where the ChlF emission is relatively high (e.g., 656, 688, and 760 nm), and the vegetation SIF can be retrieved based on spectral measurements inside and outside the Fraunhofer lines.21,22 FLD approach makes it viable to measure SIF from space, which is of significance for terrestrial SIF evaluation at large scale. While the PAM technique measures ChlF over quite a broad spectral region, SIF by FLD-based approach is estimated within very narrow spectral bands. So it requires the data of very high spectral resolution and large surface area to collect sufficient signal, which may be unavailable or costly for most remote sensors.3 Also, the atmospheric correction needs to be carefully conducted since gas scattering can result in large uncertainties to SIF retrivals.4,13 Moreover, to avoid reabsorption of fluorescence by chlorophyll, FLD-retrieved SIF is commonly measured at far-red bands, which may not be representative for PSII photochemistry analysis.4

In addition to the FLD-based methods, spectral vegetation indices (VIs) have been increasingly applied for remote estimation of ChlF-related parameters. They do not require reflectance exactly at the Fraunhofer lines but instead use reflectance at several bands sensitive to ChlF variations. These indices do not provide the direct measurements of ChlF in physical units but can be related to ChlF-derived information and parameters using empirical models.19,23,24

The methods using VIs are popular recently due to their simplicity and flexibility when applying to various remotely sensed data. Gitelson et al.25 showed that VIs with red edge reflectance were closely related to the ratio of leaf fluorescence at 735 to 700 nm; Zarco-Tejada et al.26 found a good relationship between the ratio and difference VIs derived from an airborne multispectral camera and canopy SIF estimated by FLD method; and Jia et al.27 observed that red edge position of leaf reflectance is highly correlated with Fv/Fm in winter wheat under different nitrogen treatments. These indices were mostly developed originally for estimating vegetation pigment-related bioparameters, such as chlorophyll content and leaf area index. They worked quite well for ChlF retrieval because ChlF in unstressed plants depends, to a great extent, on pigment content and leaf absorption.2830 However, it has been found that conventional chlorophyll-related VIs are not efficient to capture ChlF changes in stressed vegetation.31 Indices can still remain high when plants are heavily stressed by short-term drought while fluorescence may change dramatically.32 It is observed that vegetation chlorophyll content is not directly linked to instantaneous photosynthetic processes and cannot quantify the downregulation of photosynthesis due to early environmental stresses.7 The lag for the vegetation chlorophyll to response to short-term stress can cause significant uncertainties for ChlF estimates when using empirical algorithms developed by chlorophyll-related VIs, which have not been adequately elaborated and addressed.

This study explores using vegetation reflectance to indicate variations in ChlF-derived parameters. The experiments were conducted at leaf level since it is the basic spatial scale at which fluorescence and spectral reflectance can be measured and linked simultaneously. We selected Fv/Fm as the ChlF-derived parameter to be studied, since it reflects the potential quantum efficiency of PSII that usually acts as a sensitive indicator of plant photosynthesis performance.10 The Fv/Fm ratio has been found rather constant in nonstressed plants. A decrease in Fv/Fm can track the decrease in photochemical performance, such as sustained forms of NPQ or photoinhibition of reaction centers, when a leaf has been exposed to stress.33,34 The first objective is to test several widely used spectral VIs for Fv/Fm estimation in leaves assumed under minimum stress. The second objective is to analyze the reflectance changes as leaf under short-term stress conditions and explore the indices that are closely related to Fv/Fm. The final objective is to explore the potential of identifying leaf stress using remotely sensed reflectance data.

2.

Materials and Methods

Two experiments were conducted in this study. In experiment I, 48 fresh leaves were selected from aspen and cherry trees on the campus of Wuhan University, Hubei, China. These trees were planted under temperate climate with good management by campus services. In procedure of leaf selection, the spotted or crinkled leaves were avoided, and leaves appearing fresh and fully expanded were used. So, these leaves were assumed under minimum stress. The study period was from the early spring to late autumn in 2015, and the newly sprouted, mature, and senescent leaves were all collected. The color of sample leaves varied from yellowish-green to dark green so that they could have a wide range of chlorophyll content. The in situ measurements of reflectance, chlorophyll content, and Fv/Fm were conducted at leaf level. In experiment II, eight leaves were cut from trees and transported to the laboratory immediately. They were placed on a lab balcony having similar light and temperature conditions with outdoor environment but with no water supply. This experiment simulated short-term water stress at leaf level. Leaf reflectance, chlorophyll content, and Fv/Fm were measured every hour after the leaf was removed from the tree until Fv/Fm value was close to zero.

Reflectance spectra of the leaves were taken using a hyperspectral radiometer (ASD FieldSpec 4, ASD Inc., Boulder, Colorado) with a self-illuminated leaf clip. For each leaf, spectral reflectance in range between 400 and 900 nm at a spectral resolution of 3 nm was measured at five positions randomly distributed on the adaxial side of leaf (five scans for each position), and the median of 25 spectra scans was used as the leaf reflectance. More details can be found by Zygielbaum et al.35

Following reflectance measurements, leaf chlorophyll content was quantified using SPAD-502 (Soil and Plant Analyzer Development, Konica Minolta Inc., Ramsey, New Jersey). The SPAD can instantly provide chlorophyll meter readings based on leaf transmittance, which is widely used to assess leaf chlorophyll content.3638 Close to positions where reflectance measurements were taken, five SPAD readings were recorded and their median was calculated as leaf chlorophyll content.

Leaf Fv/Fm was measured using a PAM fluorometer (FluroPen FP 100, Photon Systems Inc., Drásov, Czech Republic). Measurements were conducted on the adaxial surface of the leaves. The leaf was first dark-adapted for 20 min using a leaf clip holder with a movable black cap. The minimum fluorescence (F0) and maximal fluorescence (Fm) were then measured and recorded by FluroPen. Fv/Fm was calculated as Fv/Fm=(FmF0)/Fm.10 More details about FluroPen were given by Aroca et al.39 and Ni et al.40

The reflectance spectra were resampled to simulate the spectral bands of Multispectral Instrument (MSI) aboard Sentinel-2.41 This satellite is showing great potential for land surface parameter estimations.42 Since the spectral response functions of the MSI spectral bands are close to rectangular,43 the simulated MSI reflectance was calculated by the average reflectance over the bandwidth of the respective MSI bands (band 1: 456 to 523 nm, band 2: 542 to 578 nm, band 3: 650 to 680 nm, band 4: 697 to 713 nm, band 5: 732 to 748 nm, and band 6: 773 to 793 nm). Using the simulated Sentinel-2 MSI reflectance, several widely used VIs were calculated (Table 1) and compared with the corresponding leaf SPAD and Fv/Fm values. The sample size for this study was 48 for the experiment I and 80 for the experiment II.

Table 1

Summary of VIs used in this study. ρ560, ρ665, ρ705, ρ740, ρ780, ρ531, and ρ570 are reflectance values simulated in Sentinel-2 bands of green, red, shortwave red edge, longwave red edge, NIR regions, 531, and 570 nm, respectively.

VIFormulaReference
NDVINDVI=(ρ780ρ665)/(ρ780+ρ665)Rouse et al.44
SRSR=ρ780/ρ665Jordan45
EVI2EVI2=2.5×(ρ780ρ665)/(ρ780+2.4×ρ665+1)Jiang et al.46
Green chlorophyll indexCIgreen=(ρ780/ρ560)1Gitelson et al.47,48
Red edge NDVINDRE705=(ρ780ρ705)/(ρ780+ρ705)Gitelson and Merzlyak49,50
NDRE740=(ρ780ρ740)/(ρ780+ρ740)
Red edge chlorophyll indexCI705=(ρ780/ρ705)1Gitelson et al.47,48
CI740=(ρ780/ρ740)1
MTCIMTCI=(ρ780ρ705)/(ρ705ρ665)Dash and Curran51
PRIPRI=(ρ531ρ570)/(ρ531+ρ570)Gamon et al.52

3.

Results and Discussion

3.1.

Remote Estimation of Chlorophyll Fluorescence in Nonstressed Leaves

In experiment I for samples assumed under minimum stress, it was observed that leaf chlorophyll varied over a wide range from SPAD values from 10 to 50, whereas Fv/Fm was rather stable in the value around 0.8±8%. Generally, Fv/Fm positively related to SPAD with the coefficient of determination (R2) above 0.76 [Fig. 1(a)]. The measured reflectances in fresh leaves at different Fv/Fm and SPAD values are presented in Fig. 1(b). With the increase in Fv/Fm and SPAD, leaf reflectance decreased in PAR region (400 to 700 nm) and increased in NIR region (750 to 900 nm). In unstressed leaves, the increase of chlorophyll content [Fig. 1(a)] results in a slight increase in Fv/Fm, and leaves with higher chlorophyll content were able to absorb more PAR thus decreasing the visible reflectance.47 In addition, leaves having higher chlorophyll content were more mature and thicker, so leaf scattering became higher causing higher NIR reflectance.

Fig. 1

(a) Relationship of relative chlorophyll content (SPAD) versus ChlF parameter (Fv/Fm). (b) Spectral reflectance of nonstressed leaves at different Fv/Fm and SPAD value.

JARS_11_2_026034_f001.png

Many studies used reflectance-derived indices to estimate leaf chlorophyll content accurately.53,54 Due to the close relationship between chlorophyll content and Fv/Fm in unstressed leaves [Fig. 1(a)], these chlorophyll-related indices may be also useful to indicate Fv/Fm variations. The relationships of Fv/Fm versus VI were tested for indices that were widely applied as good proxies of leaf chlorophyll content [Figs. 2(a)2(i)]. All chlorophyll-related VIs close to linearly related to Fv/Fm with R2 above 0.62. Since some research suggested to use photochemical reflectance index (PRI) to indicate vegetation stress,55,56 the relationship of PRI versus Fv/Fm was also tested in this study [Fig. 2(j)], but a lower correlation was found between them (R2=0.32).

Fig. 2

Relationship between Fv/Fm and (a) NDVI, (b) SR, (c) EVI2, (d) CIgreen, (e) NDRE705, (f) NDRE740, (g) CI705, (h) CI740, (i) MTCI, and (j) PRI in nonstressed leaves.

JARS_11_2_026034_f002.png

Table 2 gives the algorithms developed for Fv/Fm estimation using tested VIs. Except PRI, all indices worked well to estimate Fv/Fm with normalized root-mean-square errors (NRMSE) below 0.173 (Table 2). Note that the indices using longwave red edge band (NDRE740 and CI740) were superior to those using shortwave red edge band, and they were the most accurate for Fv/Fm retrieval among the tested indices. This is consistent with findings that the use of longwave red edge bands can give better estimates of vegetation chlorophyll content and productivity.43,57

Table 2

Best-fit functions, determination coefficients (R2), root-mean-square errors (RMSE), and NRMSE of relationships Fv/Fm versus VIs in nonstressed leaves.

VIAlgorithmR2RMSENRMSE
NDRE740Fv/Fm=1.57x+0.750.720.01250.156
CI740Fv/Fm=0.75x+0.750.720.01270.159
NDRE705Fv/Fm=0.21x+0.720.680.01290.161
CI705Fv/Fm=0.05x+0.740.670.01280.160
SRFv/Fm=0.02x+0.690.650.01300.163
MTCIFv/Fm=0.03x+0.740.650.01310.164
CIgreenFv/Fm=0.03x+0.740.640.01340.168
EVI2Fv/Fm=0.28x+0.640.640.01360.170
NDVIFv/Fm=0.27x+0.610.620.01380.173
PRIFv/Fm=1.01x+0.740.320.15991.955

3.2.

Remote Estimation of Chlorophyll Fluorescence in Stressed Leaves

In experiment II for samples assumed to be under short-term stress, it is observed that in most leaves Fv/Fm remained invariant during the first 4 or 5 h but sharply decreased afterward [e.g., Fig. 3(a)]. In experiment I, Fv/Fm values of fresh leaves ranged mostly from 0.75 to 0.85 [Fig. 1(a)], which were similar with observations by Aschan et al.,58 that this ratio differed insignificantly in unstressed plants. In experiment II, however, Fv/Fm could drop below 0.5 after 7 h of leaf removal while leaf chlorophyll content varied only slightly within a day (SPAD value almost the same). Fv/Fm indicates the intrinsic efficiency of PSII photochemistry, which is more sensitive to environmental stress.32 Under conditions of water stress, Fv/Fm declines as the PSII reaction center is damaged in the leaf.59 However, chlorophyll content does not drop immediately in response to short-term stress. Leaf-level chlorophyll changes are only visible over time scales of days,60 because plants have evolved a number of regulatory mechanisms to modulate PAR absorption before decreasing chlorophyll pigment,61 such as leaf movements, leaf angle adjustments,62 and chloroplast avoidance movements.63 Another reason may be that the chlorophyll absorption has two reaction centers (P680 for PSII and P700 for PSI), and the dissociation of them under short-term stress may not be immediately reflected in changes of bulk chlorophyll content.

Fig. 3

(a) Fv/Fm value and (b) leaf reflectance measured every hour in lab since the leaf removed from the tree with no water supply.

JARS_11_2_026034_f003.png

Figure 3(b) shows the leaf reflectance measured every hour since the leaf removed from the tree. More sensitive than chlorophyll content, leaf reflectance showed noticeable variations as stress became severe. The reflectance measured 5 h after the leaf removal was obviously different from the reflectance measured earlier, which was in agreement with Fv/Fm variations shown in Fig. 3(a) (Fv/Fm dropped also after 5 h). The changes of spectral reflectance for the green (530 to 560 nm) and far-red to NIR (700 to 900 nm) bands were the most pronounced. Before the fifth hour, reflectance reached a high and flat plateau in NIR range, and thereafter, NIR reflectance decreased but slopes in NIR and red edge regions significantly increased. In contrast, the slope in green regions slightly decreased. This is consistent with the statements by Zarco-Tejada et al.19,64 that the derivative reflectance (slope of reflectance) may enable the detection of subtle changes of ChlF emissions. We therefore analyzed the sensitivity of reflectance slope to Fv/Fm variations at different wavelengths. Since chlorophyll exhibits fluorescence emission mainly in the red and NIR regions with peaks at 690 and 740 nm,27 the first derivative of reflectance (ρ) was calculated, and the coefficients of determination (R2) between Fv/Fm and ρ were plotted with wavelength in range of 600 to 900 nm (Fig. 4). The R2 value peaked around 675 nm (above 0.8), and it remained high (between 0.80 and 0.87) throughout the wide spectral region of 760 to 860 nm. The relationship of ρ at 675 and 760 nm versus Fv/Fm was presented in Fig. 5, respectively, and it is observed that the derivative of red reflectance was positively correlated with Fv/Fm [Fig. 5(a)] while the relationship of derivative of NIR reflectance and Fv/Fm [Fig. 5(b)] was negative.

Fig. 4

The coefficient of determination (R2) between Fv/Fm and the first derivative reflectance (ρ) plotted with wavelength in range of 600 to 900 nm.

JARS_11_2_026034_f004.png

Fig. 5

Relationship between Fv/Fm and the derivative reflectance at (a) 675 and (b) 760 nm in stressed leaves.

JARS_11_2_026034_f005.png

To explore whether the VIs can be applied for Fv/Fm estimations under stress conditions, the samples of the experiment II that showed the obvious drop in Fv/Fm (usually 5 h after leaf removal with Fv/Fm below 0.75) were selected as stressed leaves. Due to the limitation of the spectral resolution of current sensors, it may be quite difficult to get hyperspectral data, thus, calculate ρ. Therefore, the indices in Table 1 were also tested in selected stressed leaves. Green chlorophyll index (CIgreen), NDRE705, and CI705, working well in nonstressed leaves (R2 above 0.64), showed very low correlations to Fv/Fm in stressed leaves with R2 below 0.28. Normalized difference vegetation index (NDVI), MERIS terrestrial chlorophyll index (MTCI), enhanced vegetation index 2 (EVI2), and simple ratio (SR), which used red band around 680 nm, showed significant relationships with Fv/Fm in stressed leaves, but R2 of these relationships in stressed leaves was much lower than in nonstressed leaves (Tables 2 and 3). R2 of the relationship PRI versus Fv/Fm increased significantly in stressed leaves (0.32 versus 0.68). Many studies showed that in ecosystems when chlorophyll content changed over a wide range, PRI was highly correlated to chlorophyll variations at both leaf and canopy level,65,66 and it may be not effective for detecting vegetation photosynthetic efficiency. But PRI acted as a useful index of photosynthetic efficiency in evergreen species when chlorophyll content changed slightly and its variation was quite independent to chlorophyll.67 Thus, when leaf chlorophyll content did not change significantly under short-term stress, PRI appeared more useful than chlorophyll content to indicate Fv/Fm variations. Among the tested VIs, CI740 and NDRE740 were the best to estimate Fv/Fm with R2 of 0.88 and NRMSE below 0.115. They were formulated based on the reflectance difference of two bands located at longwave red edge and NIR spectrum, which may indicate ρ in spectral region of 750 to 900 nm, whereas PRI may indicate ρ in green spectral region. From Fig. 3(b), it is observed that in stressed leaves, as Fv/Fm decreased, the reflectance slope in NIR the spectral region increased while it decreased in the green spectral region.

Table 3

Best-fit functions, determination coefficients (R2), RMSE, and NRMSE of relationships Fv/Fm versus VIs in stressed leaves.

VIAlgorithmR2RMSENRMSE
NDRE740Fv/Fm=7.08x+0.890.880.0710.114
CI740Fv/Fm=3.11x+0.870.880.0710.115
PRIFv/Fm=13.54x+0.360.680.1170.188
MTCIFv/Fm=0.46x+1.510.590.1340.216
NDVIFv/Fm=3.64x2.050.580.1350.218
SRFv/Fm=0.14x0.310.500.1470.237
EVI2Fv/Fm=1.47x0.120.490.1480.239
CI705Fv/Fm=0.24x+0.840.280.2050.334
NDRE705Fv/Fm=1.42x+1.090.270.2040.333
CIgreenFv/Fm=0.26x+1.180.260.1870.301

3.3.

Identify Stress Occurrence in Leaves

CI740 and NDRE740 were accurate for Fv/Fm estimates in both experiments with R2 above 0.72 (Tables 2 and 3). However, the algorithms developed in nonstressed leaves were very different from those in stressed leaves. For nonstressed leaves, CI740 and NDRE740 positively related to Fv/Fm, whereas they related to Fv/Fm negatively in stressed leaves [Figs. 6(a) and 6(b)]. Based on best-fit functions calibrated in two experiments, the stressed and nonstressed lines were roughly defined in VI versus Fv/Fm two-dimensional (2-D) space. The nonstressed line was short with relatively high Fv/Fm values. Since in healthy leaves, Fv/Fm was quite conservative with small variations. The stressed line was relatively long with higher CI740 or NDRE740 values. When leaf was under possible stresses, Fv/Fm dramatically decreased having wider dynamic range, and leaf reflectance decreased thus increasing VI values. Figures 6(c) and 6(d) show all samples of the experiment II in VI versus Fv/Fm space. Within 0 to 5 h, sample points were mostly located near the nonstressed line defined in the space. One or two hours later (the sixth and seventh hour), points moved toward the stressed line. After 8 h with no water supply, points shifted to the stressed line and quite far away from the nonstressed line. Such movements of samples in VI versus Fv/Fm space may give us a hint to identify short-term stress occurrence in leaves when leaf chlorophyll content remains stable while Fv/Fm tends to decrease.

Fig. 6

Relationship of VI versus Fv/Fm in stressed and nonstressed leaves for (a) NDRE740 and (b) CI740. Sample movements within VI versus Fv/Fm 2-D space in leaves under short-term stress for (c) NDRE740 and (d) CI740.

JARS_11_2_026034_f006.png

Our study estimated ChlF-derived Fv/Fm using spectral indices for both nonstressed and stressed leaves with promising results. The tested indices were calculated using reflectance simulated in Sentinel-2 bands, which may be easily applied to the mainstream sensors with relatively wide bands. But in this study, the experiment for stressed leaves was only conducted in lab where the light and temperature conditions might be quite different from real vegetation growing environment. Also, we studied using leaf reflectance to indicate water stress, but how the vegetation spectra will respond to other stresses (e.g., temperature and soil moisture) that may reduce photosynthesis efficiency is unclear. The next step is to apply this method for leaves in the field under various natural stresses. Furthermore, it is worthwhile to explore using reflectance-derived indices to indicate ChlF information at canopy level. Vertical gradients of light and pigment distribution inside the canopy greatly increase the complexity of ChlF signals. Also, the clarification of the relationship between fluorescence and photosynthetic efficiency at canopy level is important for monitoring ChlF remotely at large scale with aircraft or satellite data.

4.

Conclusions

We estimated a leaf ChlF-derived parameter, the intrinsic efficiency of PSII photochemistry (Fv/Fm), based on leaf spectra measurements. In leaves with minimum stress, it is found that chlorophyll content closely related to Fv/Fm. Thus, Fv/Fm can be estimated via VIs, which was developed for chlorophyll content retrieval. Several widely used indices, calculated from leaf reflectance simulated in Sentinel-2 bands, were tested to estimate Fv/Fm for nonstressed leaves. All chlorophyll-related VIs worked well, and the indices using longwave red edge bands (CI740 and NDRE740) were the most accurate for Fv/Fm retrieval with R2 above 0.72. For leaves under short-term stresses, the Fv/Fm value began to drop but leaf chlorophyll content remained almost invariant within a day. Compared to leaf chlorophyll, the first derivative of reflectance in spectra range of 700 to 900 nm appeared more sensitive to Fv/Fm variations. While most chlorophyll-related VIs showed low correlations with Fv/Fm for leaves under stress, CI740 and NDRE740 were closely related to Fv/Fm also in stressed leaves with R2 of 0.88. Note that CI740 or NDRE740 versus Fv/Fm relationships for stressed leaves were significantly different from that for nonstressed leaves, which may give a good implication to detect short-term stress occurrence in leaves.

Acknowledgments

We acknowledge the support and the use of facilities and equipment provided by the Institute of Quantitative Remote Sensing, School of Remote Sensing and Information Engineering, Wuhan University, China. This research was supported by the National Defense Project, China (No. 30-Y20A29-9003-15/17), the National 863 Project of China (No. 2013AA102401), Science Supporting Project of Hubei Province (No. 2015BCE045), and the National Natural Science Foundation of China (No. 41401390).

References

1. 

U. Schreiber, W. Bilger, C. Neubauer, “Chlorophyll fluorescence as a nonintrusive indicator for rapid assessment of in vivo photosynthesis,” Ecophysiology of Photosynthesis, 49 –70 1st ed.Springer-Verlag, Berlin, Heidelberg (1995). Google Scholar

2. 

G. H. Krause and E. Weis, “Chlorophyll fluorescence as a tool in plant physiology: II. Interpretation of fluorescence signals,” Photosynth. Res., 5 139 –157 (1984). http://dx.doi.org/10.1007/BF00028527 PHRSDI 0166-8595 Google Scholar

3. 

Z. Malenovsky et al., “Physically-based retrievals of Norway spruce canopy variables from very high spatial resolution hyperspectral data,” in IEEE Int. Geoscience and Remote Sensing Symp. (IGARSS), 4057 –4060 (2007). http://dx.doi.org/10.5167/uzh-78059 Google Scholar

4. 

A. Porcar-Castell et al., “Linking chlorophyll a fluorescence to photosynthesis for remote sensing applications: mechanisms and challenges,” J. Exp. Bot., 65 4065 –4095 (2014). http://dx.doi.org/10.1093/jxb/eru191 JEBOA6 1460-2431 Google Scholar

5. 

C. V. D. Tol, W. Verhoef and A. Rosema, “A model for chlorophyll fluorescence and photosynthesis at leaf scale,” Agric. For. Meteorol., 149 96 –105 (2009). http://dx.doi.org/10.1016/j.agrformet.2008.07.007 0168-1923 Google Scholar

6. 

C. Frankenberg et al., “New global observations of the terrestrial carbon cycle from GOSAT: patterns of plant fluorescence with gross primary productivity,” Geophys. Res. Lett., 38 351 –365 (2011). http://dx.doi.org/10.1029/2011GL048738 GPRLAJ 0094-8276 Google Scholar

7. 

Y. Zhang et al., “Estimation of vegetation photosynthetic capacity from space-based measurements of chlorophyll fluorescence for terrestrial biosphere models,” Global Change Biol., 20 3727 –3742 (2014). http://dx.doi.org/10.1111/gcb.12664 Google Scholar

8. 

U. Schreiber, “New emitter-detector-cuvette assembly for measuring modulated chlorophyll fluorescence of highly diluted suspensions in conjunction with the standard PAM fluorometer,” Z. Naturforsch. C, 49 646 –656 (1994). Google Scholar

9. 

J. E. Lee, C. Frankenberg and C. V. D. Tol, “Forest productivity and water stress in Amazonia: observations from GOSAT chlorophyll fluorescence,” Proc. R. Soc. B Biol. Sci., 280 176 –188 (2013). http://dx.doi.org/10.1098/rspb.2013.0171 Google Scholar

10. 

K. Maxwell and G. N. Johnson, “Chlorophyll fluorescence—a practical guide,” J. Exp. Bot., 51 659 –668 (2000). http://dx.doi.org/10.1093/jexbot/51.345.659 JEBOA6 1460-2431 Google Scholar

11. 

N. R. Baker, “Chlorophyll fluorescence: a probe of photosynthesis in vivo,” Annu. Rev. Plant Biol., 59 89 –113 (2008). http://dx.doi.org/10.1146/annurev.arplant.59.032607.092759 Google Scholar

12. 

A. Porcar-Castell, “A high-resolution portrait of the annual dynamics of photochemical and non-photochemical quenching in needles of Pinus sylvestris,” Physiol. Plant., 143 139 –153 (2011). http://dx.doi.org/10.1111/ppl.2011.143.issue-2 PHPLAI 0031-9317 Google Scholar

13. 

I. Moya et al., “A new instrument for passive remote sensing: 1. Measurements of sunlight-induced chlorophyll fluorescence,” Remote Sens. Environ., 91 186 –197 (2004). http://dx.doi.org/10.1016/j.rse.2004.02.012 Google Scholar

14. 

U. Rascher et al., “CEFLES2: the remote sensing component to quantify photosynthetic efficiency from the leaf to the region by measuring sun-induced fluorescence in the oxygen absorption bands,” Biogeosciences, 6 1181 –1198 (2009). http://dx.doi.org/10.5194/bgd-6-2217-2009 Google Scholar

15. 

J. Joiner et al., “First observations of global and seasonal terrestrial chlorophyll fluorescence from space,” Biogeosciences, 8 637 –651 (2011). http://dx.doi.org/10.5194/bg-8-637-2011 Google Scholar

16. 

U. Rascher, B. Gioli, F. Miglietta, “FLEX—fluorescence explorer: a remote sensing approach to quantify spatio-temporal variations of photosynthetic efficiency from space,” Photosynthesis. Energy from the Sun, 1387 –1390 Springer Netherlands, Berlin, Heidelberg, Germany (2007). Google Scholar

17. 

“Candidate Earth Explorer Core Missions—Reports for assessment: FLEX–fluorescence explorer,” Noordwijk, The Netherlands (2008). Google Scholar

18. 

C. Frankenberg et al., “Prospects for chlorophyll fluorescence remote sensing from the Orbiting Carbon Observatory-2,” Remote Sens. Environ., 147 1 –12 (2014). http://dx.doi.org/10.1016/j.rse.2014.02.007 Google Scholar

19. 

P. J. Zarco-Tejada et al., “Steady-state chlorophyll a fluorescence detection from canopy derivative reflectance and double-peak red-edge effects,” Remote Sens. Environ., 84 283 –294 (2003). http://dx.doi.org/10.1016/S0034-4257(02)00113-X Google Scholar

20. 

J. A. Plascyk, “The MK II Fraunhofer line discriminator (FLD-II) for airborne and orbital remote sensing of solar-stimulated luminescence,” Opt. Eng., 14 339 –346 (1975). http://dx.doi.org/10.1117/12.7971842 Google Scholar

21. 

L. Gomezchova et al., “Solar induced fluorescence measurements using a field spectroradiometer,” in AIP Conf. Proc., 274 –281 (2006). http://dx.doi.org/10.1063/1.2349354 Google Scholar

22. 

L. Alonso et al., “Improved Fraunhofer line discrimination method for vegetation fluorescence quantification,” IEEE Geosci. Remote Sens., 5 620 –624 (2008). http://dx.doi.org/10.1109/LGRS.2008.2001180 Google Scholar

23. 

M. Meroni et al., “Remote sensing of solar-induced chlorophyll fluorescence: review of methods and applications,” Remote Sens. Environ., 113 2037 –2051 (2009). http://dx.doi.org/10.1016/j.rse.2009.05.003 Google Scholar

24. 

S. Z. Dobrowski et al., “Simple reflectance indices track heat and water stress-induced changes in steady-state chlorophyll fluorescence at the canopy scale,” Remote Sens. Environ., 97 403 –414 (2005). http://dx.doi.org/10.1016/j.rse.2005.05.006 Google Scholar

25. 

A. A. Gitelson, C. Buschmann and H. K. Lichtenthaler, “The chlorophyll fluorescence ratio F735/F700 as an accurate measure of the chlorophyll content in plants,” Remote Sens. Environ., 69 296 –302 (1999). http://dx.doi.org/10.1016/S0034-4257(99)00023-1 Google Scholar

26. 

P. J. Zarco-Tejada et al., “Imaging chlorophyll fluorescence with an airborne narrow-band multispectral camera for vegetation stress detection,” Remote Sens. Environ., 113 1262 –1275 (2009). http://dx.doi.org/10.1016/j.rse.2009.02.016 Google Scholar

27. 

M. Jia et al., “Inversion of chlorophyll fluorescence parameters on vegetation indices at leaf scale,” in IEEE Int. Geoscience and Remote Sensing Symp. (IGARSS), 4359 –4362 (2016). Google Scholar

28. 

C. Buschmann, H. K. Lichtenthaler, “Reflectance and chlorophyll fluorescence signatures of leaves,” Applications of Chlorophyll Fluorescene in Photosynthesis Research, Stress Physiology, Hydrobiology and Remote Sensing, 325 –332 Springer Netherlands, Berlin, Heidelberg (1988). Google Scholar

29. 

H. G. Dahn, K. P. Günther and W. Lüdeker, “Characterisation of drought stress of maize and wheat canopies by means of spectral resolved laser induced fluorescence,” EARSeL Adv. Remote Sens., 1 12 –19 (1992). ERESE5 Google Scholar

30. 

J. Louis, A. Ounis and J. M. Ducruet, “Remote sensing of sunlight-induced chlorophyll fluorescence and reflectance of Scots pine in the boreal forest during spring recovery,” Remote Sens. Environ., 96 37 –48 (2005). http://dx.doi.org/10.1016/j.rse.2005.01.013 Google Scholar

31. 

G. P. Asner and A. Alencar, “Drought impacts on the Amazon forest: the remote sensing perspective,” New Phytol., 187 569 –578 (2010). http://dx.doi.org/10.1111/j.1469-8137.2010.03310.x NEPHAV 0028-646X Google Scholar

32. 

S. Wang et al., “Monitoring and assessing the 2012 drought in the Great Plains: analyzing satellite-retrieved solar-induced chlorophyll fluorescence, drought indices, and gross primary production,” Remote Sens., 8 (2), 61 (2016). http://dx.doi.org/10.3390/rs8020061 Google Scholar

33. 

C. Ottander and G. Oquist, “Recovery of photosynthesis in winter-stressed Scots pine,” Plant Cell Environ., 14 345 –349 (1991). http://dx.doi.org/10.1111/pce.1991.14.issue-3 Google Scholar

34. 

A. Porcar-Castell et al., “Seasonal acclimation of photosystem II in Pinus sylvestris. II. Using the rate constants of sustained thermal energy dissipation and photochemistry to study the effect of the light environment,” Tree Physiol., 28 1483 –1491 (2008). http://dx.doi.org/10.1093/treephys/28.10.1483 TRPHEM 0829-318X Google Scholar

35. 

A. I. Zygielbaum et al., “Non‐destructive detection of water stress and estimation of relative water content in maize,” Geophys. Res. Lett., 36 91 –100 (2009). http://dx.doi.org/10.1029/2009GL038906 GPRLAJ 0094-8276 Google Scholar

36. 

C. W. Wood et al., “Field chlorophyll measurements for evaluation of corn nitrogen status,” J. Plant Nutr., 15 487 –500 (1992). http://dx.doi.org/10.1080/01904169209364335 Google Scholar

37. 

C. W. Wood, D. W. Reeves and D. G. Himelrick, “Relationships between chlorophyll meter readings and leaf chlorophyll concentration, N status, and crop yield: a review,” in Proc. of Agronomy Society New Zealand, 1 –9 (1993). Google Scholar

38. 

Q. Ling, W. Huang and P. Jarvis, “Use of a SPAD-502 meter to measure leaf chlorophyll concentration in Arabidopsis thaliana,” Photosynth. Res., 107 209 –214 (2011). http://dx.doi.org/10.1007/s11120-010-9606-0 PHRSDI 0166-8595 Google Scholar

39. 

R. Aroca et al., “Arbuscular mycorrhizal symbiosis influences strigolactone production under salinity and alleviates salt stress in lettuce plants,” J. Plant Physiol., 170 (1), 47 –55 (2013). http://dx.doi.org/10.1016/j.jplph.2012.08.020 JPPHEY 0176-1617 Google Scholar

40. 

L. Ni et al., “Effects of artemisinin on photosystem II performance of microcystis aeruginosa by in vivo chlorophyll fluorescence,” Bull. Environ. Contam. Toxicol., 89 1165 –1169 (2012). http://dx.doi.org/10.1007/s00128-012-0843-0 Google Scholar

41. 

M. Drusch et al., “Sentinel-2: ESA’s optical high-resolution mission for GMES operational services,” Remote Sens. Environ., 120 25 –36 (2012). http://dx.doi.org/10.1016/j.rse.2011.11.026 Google Scholar

42. 

C. Atzberger and K. Richter, “Spatially constrained inversion of radiative transfer models for improved LAI mapping from future Sentinel-2 imagery,” Remote Sens. Envion., 120 208 –218 (2012). http://dx.doi.org/10.1016/j.rse.2011.10.035 Google Scholar

43. 

J. G. Clevers and A. A. Gitelson, “Remote estimation of crop and grass chlorophyll and nitrogen content using red-edge bands on Sentinel-2 and-3,” Int. J. Appl. Earth Obs. Geoinf., 23 344 –351 (2013). http://dx.doi.org/10.1016/j.jag.2012.10.008 Google Scholar

44. 

J. W. Rouse et al., “Monitoring vegetation systems in the great plains with ERTS,” in Proc. of the Third Earth Resources Technology Satellite–1 Symp., 309 –317 (1974). Google Scholar

45. 

C. F. Jordan, “Derivation of leaf-area index from quality of light on the forest floor,” Ecology, 50 663 –666 (1969). http://dx.doi.org/10.2307/1936256 ECGYAQ 0094-6621 Google Scholar

46. 

Z. Jiang et al., “Development of a two-band enhanced vegetation index without a blue band,” Remote Sens. Environ., 112 3833 –3845 (2008). http://dx.doi.org/10.1016/j.rse.2008.06.006 Google Scholar

47. 

A. A. Gitelson, Y. Gritz and M. N. Merzlyak, “Relationships between leaf chlorophyll content and spectral reflectance and algorithms for non-destructive chlorophyll assessment in higher plant leaves,” J. Plant. Physiol., 160 271 –282 (2003). http://dx.doi.org/10.1078/0176-1617-00887 JPPHEY 0176-1617 Google Scholar

48. 

A. A. Gitelson et al., “Remote estimation of canopy chlorophyll content in crops,” Geophys. Res. Lett., 32 93 –114 (2005). http://dx.doi.org/10.1029/2005GL022688 GPRLAJ 0094-8276 Google Scholar

49. 

A. A. Gitelson and M. N. Merzlyak, “Quantitative estimation of chlorophyll-a using reflectance spectrum: experiments with autumn chestnut and maple leaves,” J. Photochem. Photobiol. B, 22 247 –252 (1994). http://dx.doi.org/10.1016/1011-1344(93)06963-4 JPPBEG 1011-1344 Google Scholar

50. 

A. A. Gitelson and M. N. Merzlyak, “Remote estimation of chlorophyll content in higher plant leaves,” Int. J. Remote Sens., 18 (12), 2691 –2697 (1997). http://dx.doi.org/10.1080/014311697217558 IJSEDK 0143-1161 Google Scholar

51. 

J. Dash and J. Curran, “The MERIS terrestrial chlorophyll index,” Int. J. Remote Sens., 25 5403 –5413 (2004). http://dx.doi.org/10.1080/0143116042000274015 IJSEDK 0143-1161 Google Scholar

52. 

J. A. Gamon, J. Penuelas and C. B. Field, “A narrow waveband spectral index that tracks diurnal changes in photosynthetic efficiency,” Remote Sens. Environ., 41 35 –44 (1992). http://dx.doi.org/10.1016/0034-4257(92)90059-S Google Scholar

53. 

C. S. T. Daughtry et al., “Estimating corn leaf chlorophyll concentration from leaf and canopy reflectance,” Remote Sens. Environ., 74 229 –239 (2000). http://dx.doi.org/10.1016/S0034-4257(00)00113-9 Google Scholar

54. 

D. A. Sims and J. A. Gamon, “Relationships between leaf pigment content and spectral reflectance across a wide range of species, leaf structures and developmental stages,” Remote Sens. Environ., 81 337 –354 (2002). http://dx.doi.org/10.1016/S0034-4257(02)00010-X Google Scholar

55. 

F. Thenot, M. Méthy and T. Winkel, “The photochemical reflectance index (PRI) as a water-stress index,” Int. J. Remote Sens., 23 5135 –5139 (2002). http://dx.doi.org/10.1080/01431160210163100 IJSEDK 0143-1161 Google Scholar

56. 

M. Meroni et al., “Assessing steady-state fluorescence and PRI from hyperspectral proximal sensing as early indicators of plant stress: the case of ozone exposure,” Sensors, 8 1740 –1754 (2008). http://dx.doi.org/10.3390/s8031740 SNSRES 0746-9462 Google Scholar

57. 

Y. Peng and A. Gitelson, “Remote estimation of gross primary productivity in soybean and maize based on total crop chlorophyll content,” Remote Sens. Environ., 117 440 –448 (2012). http://dx.doi.org/10.1016/j.rse.2011.10.021 Google Scholar

58. 

G. Aschan et al., “Photosynthetic performance of vegetative and reproductive structures of green hellebore (Helleborus viridis L. agg.),” Photosynthetica, 43 55 –64 (2005). http://dx.doi.org/10.1007/s11099-005-5064-x PHSYB5 0300-3604 Google Scholar

59. 

Y. J. Zhang et al., “Chlorophyll fluorescence detected passively by difference reflectance spectra of wheat (Triticum aestivum L.) leaf,” J. Integr. Plant Biol., 47 (10), 1228 –1235 (2005). http://dx.doi.org/10.1111/jipb.2005.47.issue-10 Google Scholar

60. 

J. I. García-Plazaola and J. M. Becerril, “Seasonal changes in photosynthetic pigments and antioxidants in beech (Fagus sylvatica) in a Mediterranean climate: implications for tree decline diagnosis,” Funct. Plant Biol., 28 225 –232 (2001). http://dx.doi.org/10.1071/PP00119 Google Scholar

61. 

R. G. Walters, “Towards an understanding of photosynthetic acclimation,” J. Exp. Bot., 56 435 –447 (2005). http://dx.doi.org/10.1093/jxb/eri060 JEBOA6 1460-2431 Google Scholar

62. 

F. Yu and V. S. Berg, “Control of paraheliotropism in two phaseolus species,” Plant Physiol., 106 1567 –1573 (1994). http://dx.doi.org/10.1104/pp.106.4.1567 Google Scholar

63. 

M. Kasahara et al., “Chloroplast avoidance movement reduces photodamage in plants,” Nature, 420 829 –832 (2002). http://dx.doi.org/10.1038/nature01213 Google Scholar

64. 

P. J. Zarco-Tejada et al., “Chlorophyll fluorescence effects on vegetation apparent reflectance: I. Leaf-level measurements and model simulation,” Remote Sens. Environ., 74 582 –595 (2000). http://dx.doi.org/10.1016/S0034-4257(00)00148-6 Google Scholar

65. 

A. A. Gitelson, J. A. Gamon and A. Solovchenko, “Mulitiple drivers of seasonal change in PRI: implications for photosynthesis 1. Leaf level,” Remote Sens. Environ., 191 110 –116 (2017). http://dx.doi.org/10.1016/j.rse.2016.12.014 Google Scholar

66. 

A. A. Gitelson, J. A. Gamon and A. Solovchenko, “Multiple drivers of seasonal change in PRI: implications for photosynthesis 2. Stand level,” Remote Sens. Environ., 190 198 –206 (2017). http://dx.doi.org/10.1016/j.rse.2016.12.015 Google Scholar

67. 

Y. Peng et al., “Remote estimation of gross primary production in maize and support for new paradigm based on total crop chlorophyll content,” Remote Sens. Environ., 115 978 –989 (2011). http://dx.doi.org/10.1016/j.rse.2010.12.001 Google Scholar

Biography

Yi Peng received her PhD from the School of Natural Resources, University of Nebraska–Lincoln, USA, in 2012. Currently, she is the associate professor in the School of Remote Sensing and Information Engineering, Wuhan University, China. Her research interests focus on quantitative remote sensing of agriculture.

Shenghui Fang is the director of the Institute of Quantitative Remote Sensing, School of Remote Sensing and Information Engineering, Wuhan University, China. His research interests include quantitative remote sensing, image processing and microwave remote sensing.

Yan Gong works as the associate director of the Institute of Quantitative Remote Sensing, School of Remote Sensing and Information Engineering, Wuhan University, China. His research interests include remote sensing of agriculture and image analysis.

Kan Liu received his PhD form the Institute of Pattern Recognition and Artificial Intelligence, Huazhong University of Science and Technology, China, in 2012. He is currently working in Wuhan Institute of Physics and Mathematics of Chinese Academy of Sciences, China. His research interests include bioinformatics, fluorescent data analysis, and integrated spectral imaging devices.

Biographies for the other authors are not available.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 Unported License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Yi Peng, Aoli Zeng, Tinge Zhu, Shenghui Fang, Yan Gong, Yanqi Tao, Ying Zhou, and Kan Liu "Using remotely sensed spectral reflectance to indicate leaf photosynthetic efficiency derived from active fluorescence measurements," Journal of Applied Remote Sensing 11(2), 026034 (19 June 2017). https://doi.org/10.1117/1.JRS.11.026034
Received: 4 January 2017; Accepted: 2 June 2017; Published: 19 June 2017
Lens.org Logo
CITATIONS
Cited by 17 scholarly publications.
Advertisement
Advertisement
KEYWORDS
Reflectivity

Luminescence

Near infrared

Infrared radiation

Photochemistry

Photosynthesis

Radiometry

Back to Top