Next Article in Journal
Genome and Transcriptome Analyses of Genes Involved in Ascorbate Biosynthesis in Pepper Indicate Key Genes Related to Fruit Development, Stresses, and Phytohormone Exposures
Next Article in Special Issue
An Inverse Scaling Relationship between Stomatal Density and Mean Nearest Neighbor Distance: Evidence from a Photinia Hybrid and One of Its Parents
Previous Article in Journal
Alternative Tissue Sampling for Improved Detection of Candidatus Liberibacter asiaticus
Previous Article in Special Issue
Inequality Measure of Leaf Area Distribution for a Drought-Tolerant Landscape Plant
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Relationship between Photosynthetic CO2 Assimilation and Chlorophyll Fluorescence for Winter Wheat under Water Stress

1
College of Natural Resources and Environment, Northwest A&F University, Xianyang 712100, China
2
State Key Laboratory of Soil Erosion and Dryland Farming on the Loess Plateau, Northwest A&F University, Xianyang 712100, China
3
The Research Center of Soil and Water Conservation and Ecological Environment, Chinese Academy of Sciences and Ministry of Education, Xianyang 712100, China
4
Institute of Soil and Water Conservation, Chinese Academy of Sciences and Ministry of Water Resources, Xianyang 712100, China
5
College of Resources and Environment, University of Chinese Academy of Sciences, Beijing 100049, China
*
Author to whom correspondence should be addressed.
Plants 2023, 12(19), 3365; https://doi.org/10.3390/plants12193365
Submission received: 30 August 2023 / Revised: 17 September 2023 / Accepted: 21 September 2023 / Published: 23 September 2023
(This article belongs to the Special Issue Multiple Response Mechanisms of Plants to Drought Stress)

Abstract

:
Solar-induced chlorophyll fluorescence (SIF) has a high correlation with Gross Primary Production (GPP). However, studies focusing on the impact of drought on the SIF-GPP relationship have had mixed results at various scales, and the mechanisms controlling the dynamics between photosynthesis and fluorescence emission under water stress are not well understood. We developed a leaf-scale measurement system to perform concurrent measurements of active and passive fluorescence, and gas-exchange rates for winter wheat experiencing a one-month progressive drought. Our results confirmed that: (1) shifts in light energy allocation towards decreasing photochemistry (the quantum yields of photochemical quenching in PSII decreased from 0.42 to 0.21 under intermediate light conditions) and increasing fluorescence emissions (the quantum yields of fluorescence increased to 0.062 from 0.024) as drought progressed enhance the degree of nonlinearity of the SIF-GPP relationship, and (2) SIF alone has a limited capacity to track changes in the photosynthetic status of plants under drought conditions. However, by incorporating the water stress factor into a SIF-based mechanistic photosynthesis model, we show that drought-induced variations in a variety of key photosynthetic parameters, including stomatal conductance and photosynthetic CO2 assimilation, can be accurately estimated using measurements of SIF, photosynthetically active radiation, air temperature, and soil moisture as inputs. Our findings provide the experimental and theoretical foundations necessary for employing SIF mechanistically to estimate plant photosynthetic activity during periods of drought stress.

1. Introduction

Light energy absorbed by plants is consumed in three competing pathways: photochemistry (photochemical quenching (PQ)), emission of chlorophyll a fluorescence (ChlF), and non-photochemical quenching (NPQ) [1]. ChlF emission is the radiative loss of absorbed solar energy in the spectral range from 640 nm to 850 nm, with emission peaks at 685 nm and 740 nm [2]. NPQ is the process by which plants dissipate absorbed photon energy as heat and consists of two components: basal or constitutive heat dissipation (D), and regulated heat dissipation (N). Energy partitioning between these three pathways may be highly dynamic under changing physiological and environmental conditions [3,4]. Thus, by measuring ChlF, one may obtain valuable information on the other two processes, namely, PQ and NPQ. Active ChlF measurements, mainly based on the active pulse-amplitude modulation (PAM) technique [5], have been successfully used to assess changes in photosynthetic machinery and the photosynthetic status of plants [6,7]. However, they are generally applied at the leaf scale because active manipulation of the light environment is required [3]. In recent years, substantial progress has also been made in the passive detection of solar-induced chlorophyll fluorescence (SIF) within solar or telluric absorption features (e.g., O2-A at 760 nm), enabling the top-of-canopy (TOC) SIF observations (SIFtoc) in discrete and narrow wavelength bands to be obtained from remote sensing platforms [8,9,10]. Many studies have shown that canopy SIF has a strong capacity to predict gross primary productivity (GPP) across a variety of land cover types [4,11,12,13]. Previous studies have indicated that GPP and SIF have a strong linear relationship, and usually exhibit similar spatial and temporal patterns [9,14,15]. Marrs et al. (2020) [16] suggested that the linear relationship between SIF and GPP at large spatial and temporal scales is the result of a shared driver.
As a consequence of climate change, droughts are expected to increase in frequency, duration and severity in many parts of the world, most notably in Africa, Asia and Central and South America [17], and the drought-induced reduction in crop yields has received widespread attention [18,19,20,21]. It is noteworthy that winter wheat is the one of the most important and widely planted staple crops in the world [22], and the normal growth and development of winter wheat can be influenced by drought stress [23,24,25]. Therefore, a rigorous evaluation of the performance of SIF in detecting early signs of photosynthetic downregulation during droughts is particularly relevant to food security. However, SIF-based studies in this field have reported inconsistent results: satellite or near-ground SIF measurements show remarkable declines during droughts [26,27,28], but leaf-level SIF/ChlF measurements have a weak and delayed response to water stress [29,30]. Moreover, a key step in estimating GPP from remotely sensed SIF is to parameterize the SIF-GPP relationship with data-driven statistical approaches: a combination of flux-tower measurements and satellite SIF data are used. These approaches do not usually track photosynthesis at regional or global scales, particularly facing the complexity of naturally varying systems. All these controversies and limitations highlight the urgent need to develop a practical approach for estimating GPP. This approach should be based on a mechanistic understanding of the relationship between ChlF and photosynthetic CO2 assimilation under drought conditions.
To obtain a complete picture of the mechanisms regulating ChlF/SIF emission and photosynthetic CO2 assimilation under a water deficit, one should not only measure passive ChlF radiance for drought plants, but also, simultaneously, gas exchange and active PAM fluorescence. In this study, we developed a concurrent leaf-level measurement system consisting of a portable gas-exchange system, a PAM instrument, and four high-resolution spectrometers. Using this measurement system, we obtained a variety of key photosynthetic parameters, actively and passively induced ChlF, and the light–response curves of gas exchange for both drought-affected and control plants during a month-long progressive drought experiment. We first show the cascade of decline in these photosynthetic parameters, which include stomatal conductance, net photosynthetic carbon assimilation, electron transport rate, and ChlF emission, in response to water stress. We present the variations in the quantum yields of photochemical quenching in PSII (ФP), fluorescence (ФF), and regulated and basal heat dissipation (ФN and ФD), namely, the probability of an absorbed photon being used in a given pathway, under different drought and light conditions. We pay particular attention to the response of the phase-shift in the ФFP relationship to water stress and its implications for interpreting the SIF-GPP relationship. By reformulating the mechanistic light–response (MLR) model [31], we are able to propose a SIF-based mechanistic model to accurately track rapid changes in plant photosynthesis status resulting from drought stress. We discuss (1) the mechanisms regulating the interaction among ФP, ФF, ФN, and ФD under different degrees of water stress, (2) the reason why the drought response of satellite SIF observations is more pronounced than that obtained from leaf-level measurements, and (3) practical considerations regarding the application of the proposed model to large scales.

2. Results

2.1. The Cascade of Drought-Induced Changes in the Photosynthetic Parameters

The morphology of the plants in the WS treatment significantly changed in response to progressive drought stress over the time course of 1 to 28 days (WS1 to WS28), resulting in the curling (WS7 and 10), yellowing (WS14), drying (WS20) and, ultimately, the death of stems/leaves (WS26) (Figure 1). The θSWC remained at 19.1% in the control pots but declined in drought-treated pots as the drought intensified. The θSWC reduction rate was high during the early stages of drought, decreasing by nearly half from 19.1% on WS0 (no stress) to 9.2% on WS9 (Figure 1), with the θSWC decreasing at a slower rate to 7.3% on day 16 of the drought cycle (WS16) as the drought stress progressed (Figure 1). Subsequently, after the imposition of progressive drought for 18 days (WS18), the θSWC remained rather stable, only declining by 6.3% by WS28 (Figure 1). The variations in βS and βB as the soil dried are provided in the Supporting Information (Figure S1).
Changes in the light–response curves of Anet, GS, Ja_PAM, and ChlFP_F with progressive drought are shown in Figure 2. A strong drought response was observed in Anet. Increasing water stress diminished both the maximum Anet value achieved and the irradiance at which this maximum was observed (Figure 2, first row). Light-saturated Anet decreased from 24.3 μmol m−2 s−1 at day 0 (prior to drought treatment) to 19.1 μmol m−2 s−1 at day 5, during which time SWC dropped from 19.1% to 12%. Anet decreased at a greater magnitude subsequently. Anet in the water-stressed plants fell to near-zero by day ten of the drought. Compared with Anet, GS showed an even more pronounced drought response (Figure 2, second row). Light-saturated GS dropped by more than 60%, or from 0.54 mol m−2 s−1 to 0.21 mol m−2 s−1, after only 5 days of drought, and the stomata almost completely closed on the 8th day of treatment (i.e., GS ≈ 0 during each light regime). Drought stress had a relatively smaller effect on the electron transport rate (Figure 2, third row). The maximum Ja_PAM reached 120.2 μmol m−2 s−1 after 5 days of water stress, or about 85% of the corresponding value at the beginning of the experiment. Over 10 days of drought, the maximum Ja_PAM still accounted for about 30% of that at WS0, while both Anet and GS approached 0. The fluorescence response occurred much later, and was much smaller, than the responses of Anet, GS, and Ja_PAM (Figure 2, bottom row); there were only small variations in the light–response curve of ChlFP_F during the first 12 days into the drought. We found that absorbed PAR in the water-stressed plants remained almost unchanged during this 12-day period of water stress development (Figure S2), which may explain the weak drought response in the fluorescence emission. In fact, the ChlFP_F value of the water-stressed plants maintained a fairly high level of ChlFP_F after two weeks of water stress, and was still detectable at WS26 (Figure S3).

2.2. Drought-Induced Changes in the Photosynthesis–Fluorescence Relationship

We compared changes in ChlFP_F during the light–response curves against the Anet of individual samples in the WS treatment under different water stress conditions: zero (θSWC = 19.1%), moderate (θSWC = 12.0%), and high (θSWC = 9.7%), which occurred at WS0, WS5, and WS8, respectively (Figure 3). Water stress regulated both the saturation levels of Anet and how the saturation level was approached as ChlFP_F increased. Under no drought stress, Anet showed an initial linear increase with increasing ChlFP_F and reached around 25 μmol m−2 s−1 when ChlFP_F = 10 μmol m−2 s−1, after which it largely leveled off with a further increase in ChlFP_F (Figure 3). As water stress increased, Anet tended to increase less steeply and reached a plateau at a lower ChlFP_F. For the plants exposed to moderate drought, for example, Anet started to remain stable at nearly 17 μmol m−2 s−1 when ChlFP_F > 6 μmol m−2 s−1 (Figure 3). At severe water stress levels, the dynamic of ChlFP_F was relatively less affected; ChlFP_F still increased from 0 to 20 μmol m−2 s−1, with PAR ranging from 0 to 2100 μmol m−2 s−1. However, Anet started to saturate even when ChlFP_F > 4 μmol m−2 s−1 and remained less than 8 μmol m−2 s−1 for the complete light–response curve (Figure 3).

2.3. The Variations in the Mechanisms Linking Photosynthesis and Fluorescence under Drought

The increasing nonlinearity in the relationship between Anet and ChlFP_F (Figure 3) suggests that water stress causes shifts in the allocation of absorbed light energy dissipation pathways. At the beginning of the experiment (WS0, non-stress), ФP showed an inverse correlation with PAR and exhibited less sensitivity to PAR with increased light levels (Figure 4). The trade-offs between these yields appear to be governed by the complementarity between PQ and NPQ: ФN increased with increased PAR, and ФP + ФN ≈ 0.8 (Figure 4). In contrast, ФD and ФF had a muted sensitivity to changes in PAR: both of them showed a slight increasing trend at low light levels, and a decreasing trend at intermediate or high light levels (Figure 4).
During the first five days after imposing water treatment (WS1 to WS5), when θSWC decreased from 19.1% to 12.0%, the light–response curves of ФP, ФN, ФD and ФF showed a weak response to drought under low PAR conditions. For example, ФP taken at PAR = 180 μmol m−2 s−1 decreased slightly from 0.63 to 0.61. However, a further reduction in ФP occurred at higher PAR levels; for instance, ФP at PAR = 700 μmol m−2 s−1 decreased from 0.42 to 0.36 (Figure 4). ФN showed a small increasing trend during each light regime, with a larger magnitude under intermediate and high light conditions (Figure 4). Both ФD and ФF increased over the entire light–response curve (Figure 4).
From the 6th to the 8th day of treatment (WS6 to WS8), the variations in these four yields were still relatively limited when PAR ≤ 700 μmol m−2 s−1. However, ФP showed a significant decline at high PAR values: at WS8, ФP decreased to 0.13, 0.10, and 0.07 at 1300, 1700, and 2100 μmol m−2 s−1, respectively (Figure 4). NPQ still remained complementary to PQ: ФNPQ increased to 0.51, 0.62, 0.65 for these three PAR regimes (Figure 4). Their net effect on ФD and ФF was diminished; the light responses of ФD remained fairly unchanged and ФF showed a small decreasing trend (Figure 4).
During the period between 9 and 12 days (WS9 to WS12), after withholding water, when θSWC dropped below 9.0%, ФP decreased rapidly with increasing PAR, and this drop grew steeper as water stress developed (Figure 4). In other words, additional water stress increased the degree of nonlinearity in the relationship between ФP and PAR. We also found that a severe water deficit may reduce the complementarity between PQ and NPQ; ФN also showed a clear decreasing trend throughout the light–response curves (Figure 4). As these four pathways compete for absorbed energy, both ФD and ФF showed a clear increase: the maximum ФD and ФF reached 0.57 and 0.03, respectively, on the 12th day of treatment (Figure 4).

2.4. The Phase-Shift in the Relationship between Photochemical and Fluorescence Yields

The nonlinear relationship between ФF and ФP physiologically regulates the asymptotic behavior of the link between fluorescence and photosynthesis, that is, a positive or negative SIF-GPP relationship [32,33]. Similarly to Maguire et al. [34], we fitted a polynomial model to the relationship between ФF and ФP, and the breakpoint was identified as the value of ФP where the slope of polynomial shifted from positive to negative (Figure 5). This breakpoint separates the relationship between ФF and ФP into two parts [3]: (1) ФF is proportional to ФP under low ФP (i.e., high light, ‘NPQ phase’), and (2) ФF is inversely proportional to ФP under high ФP (i.e., low light, ‘PQ phase’).
At WS0, the breakpoint was located at ФP = 0.46. ФF was positively correlated with ФP when ФP ≤ 0.46, and they were negatively correlated when ФP > 0.46 (Figure 5). The imposition of progressive drought for 5 days (WS1 to WS5) made the phase-shift in the ФPF relationship occur at a lower ФP; the value of ФP at the breakpoints dropped from 0.46 to 0.37 (Figure 5). However, the breakpoints showed no clear trend during the period from WS6 to WS9 (Figure 5), most likely due to a decrease in ФF during that period (i.e., a flatter relationship between ФP and ФF), and the limited number of light regimes in the middle of the light–response curves. The breakpoints were again observed to decline markedly after 10 days of progressive water stress. For example, the breakpoint at WS12 occurred at ФP = 0.12 (Figure 5).

2.5. The Performance of the rMLR Model

The trained parameters were applied to the testing dataset, and the resulting model performance in simulating Vcmax, Jmax, ФP, NPQ Anet, and GS was quantified using linear regression analysis and described using the coefficient of determination (R2) and the root mean squared error (RMSE) between the simulated and measured values.
A relatively small decrease was observed in both Vcmax and Jmax between WS1 and WS6 (Figure 6). As the drought continued, however, they dropped substantially, and were almost zero after WS10 (Figure 6). As water stress progressed, the rMLR model was able to track the decreasing trends in Vcmax, and Jmax well, explaining 83% (RMSE = 8.25 μmol m−2 s−1, Figure 6a) and 79% (RMSE = 24.35%, Figure 6b) in their variance between WS1 and WS12. However, the model tended to underestimate the large values of Vcmax, and Jmax during WS1-WS6, and overestimate the small values between WS7 and WS9 (Figure 6).
During the first week of the experiment, ФP gradually declined from 1.0 to approximately 0.3 as PAR increased from 0 to 1000 μmol m−2 s−1 (Figure 7). As the drought intensified, ФP became less responsive to PAR: ФP decreased more steeply with increased PAR and leveled off over a broader range of light intensities (Figure 7). The simulated ФP light-curve shapes were similar to those estimated from the fluorescence parameters: between WS1 and WS12, the model explained 94.3% of the variation in ФP (RMSE = 0.07, Figure 7). Under severe water stress conditions (WS8–WS12), the model underestimated ФP, with R2 = 0.93 (RMSE = 0.08, Figure 7).
The NPQ light–response curves have an approximately opposite form to those of ФP. NPQ tended to saturate at lower PAR levels with an increase in the water deficit, and substantially decrease under severe stress (Figure 8). The proposed model was able to track the light response of NPQ well: R2 between measured and modelled NPQ reached a value of 0.97 (RMSE = 0.16) for the period between WS0 and WS12 (Figure 8). The decrease in θSWC had no obvious effect on the performance of the model in simulating NPQ, with R2 remaining at 0.97 between WS9 and WS12 (RMSE = 0.16, Figure 8). Despite the overall good performance, a small underestimation in NPQ was observed under high light intensities (PAR > 1700 μmol m−2 s−1, Figure 8).
The simulated response of Anet to variations in light intensity shows a similar pattern to the response measured by the gas-exchange system: simulated Anet increases rapidly with increasing illumination intensity at low light levels, and gradually reaches a plateau under high light conditions (Figure 9). The rMLR model is able to reproduce Anet well under drought conditions; it accounted for 97.2% (RMSE = 1.532 μmol m−2 s−1) of the variability in Anet from WS0 to WS12 (Figure 9). However, the model did tend to consistently overestimate Anet, with the degree of overestimation appearing to be independent of the severity of drought; simulated Anet was about 15% higher than measured Anet in non-stressed plants and in plants subjected to mild drought (Figure 9), with the extent of overestimation varying between 10% and nearly 50% for moderate and severe drought, respectively (Figure 9). The constant value used for the ФPmax value (0.8) is one possible explanation for the overestimation. A decreased ФPmax has been suggested to occur under water stress [1,35,36]. An overestimation of ФPmax in drought would lead to an overestimation in χ (Equation (12)) and then in NPQ (Equation (10)), and would consequently lead to a higher simulated Anet than that observed (NPQ occurs in the numerator of the rMLR model, Equation (3)).
Overall, the simulated GS can explain 91.7% (RMSE = 0.044 μmol m−2 s−1) of the variability in the measurements of GS collected between WS0 and WS12 (Figure 10). The rMLR model overestimated GS for non-stressed plants (WS0) by about 15%. However, the overestimation was suppressed from the beginning of the drought-stress period: the RMSE between simulated and measured GS decreased to 0.040 μmol m−2 s−1 (WS1–WS12) for the water-stressed plants (Figure 10). In particular, the proposed model demonstrated good potential for mimicking the fast drop in stomatal conductance (R2 = 0.89 and RMSE = 0.035 μmol m−2 s−1) in the early stages of water stress (WS1–WS4). The overestimation of GS at WS0 can be explained by the overestimation in Anet (Figure 9). The decrease in βS as drought became more severe tended to cancel out the negative effect of overestimation in Anet (Equation (21)), resulting in an improvement in the simulation of GS.

3. Discussion

Identification of the cascade of drought-induced changes in photosynthetic characteristics is highly relevant to fully understanding the relationships between photosynthesis, NPQ, and fluorescence under water stress conditions. Stomatal closure is the earliest response to drought, paralleling a decrease in photosynthetic CO2 assimilation but with a lower magnitude (Figure 2). In the early stages of drought stress, exposure to the volatile plant hormone methyl jasmonate (MeJA) can mitigate drought stress by stomatal closure [37,38,39] During WS1–WS6, under mild-to-moderate water stress, energy partitioning in PSII (i.e., ФP, ФN, ФD, and ФF) remained virtually unchanged under low or intermediate light levels (Figure 4). However, we noticed a decline in ФP, and a rise in the other three pathways, ФD, ФN and ФF, in the presence of high light levels during that period (Figure 4). These observations may suggest that the proportion of energy dissipated in each pathway does not change much regardless of stomatal closure levels during the initial period of drought [40].
At moderate-to-severe drought (WS7–WS9), GS almost vanished (Figure 2) and the observed drop in Anet indicated the predominance of non-stomatal limitations to photosynthesis [41]. Given the lower GS, the excess light energy can ultimately lead to the production of reactive oxygen species and reflect the MeJA response [37,38,39]. As a consequence, ФP, the quantum yield of photochemistry in PSII, showed a decreasing trend, especially at high light levels (Figure 4). To compensate for the decline in ФP, regulated heat dissipation (ФN) tended to increase as a main photoprotective mechanism, and ФD remained fairly constant (Figure 4). The engagement of thermal energy dissipation also resulted in a small decrease in ФF.
Under extreme drought conditions (WS10–WS12), ФP continued to decrease, and the stressed plants kept increasing their use of thermal energy dissipation and fluorescence emission (ФF) to cope with excess light energy (Figure 4). However, the process of thermal energy dissipation changed from a regulated form that can be rapidly activated by excess light to a sustained form that is rather insensitive to fluctuating light [42,43], resulting in the simultaneous rise of ФD and decline of ФN (Figure 4). This result is consistent with the findings that this transformation of energy dissipation characteristics tends to occur during periods of harsh environmental stress [44].
The results confirm that water stress increases nonlinearity in the overall relationship between photosynthesis and fluorescence (Figure 3). Under typical high light levels, as what might occur for early afternoon spaceborne SIF retrievals, the ‘NPQ phase’, driving a positive SIF-GPP relationship, is representative of non-stressed plants. We showed that the transition from the ‘PQ phase’ to the ‘NPQ phase’ is driven by both irradiance and the severity of drought. During moderate or severe drought stress, the decrease in ФP, and increase in ФF, pushes the stressed plant from the ‘NPQ phase’ to the ‘PQ phase’, resulting in a more nonlinear SIF-GPP relationship when considering both the drought and non-drought periods together.
When the stomata are closed, Marrs et al. (2020) [16] found that the photosynthetic rate decreases rapidly, and SIF is relatively less affected. SIF cannot track the changes of photosynthesis, resulting in the phenomenon of decoupling fluorescence and photosynthesis. During a short drought duration, Helm et al. (2020) [29] showed that the degree of reduction in leaf photosynthesis was much greater than the reduction in SIF. These studies are consistent with our findings. Our results also show that the fluorescence response to water stress occurs later, and at a smaller magnitude, which appears to contradict studies using satellite SIF, suggesting a strong negative response [26,27]. Such a dichotomy can be resolved by assessing the relative importance of the structural and physiological contributions to the drought response of fluorescence at different scales. Fluorescence variations reported in leaf-scale studies are driven by changing fluorescence efficiencies alone (unless absorbed PAR is unchanged) which, as shown here, has a relatively muted sensitivity to water stress. However, at the whole-plant or canopy scales, both structural and physiologic components may regulate TOC SIF, and so a drought-induced decline in absorbed PAR due to canopy structural changes (i.e., changes in leaf area or leaf angles) dominates negative anomalies observed in satellite SIF [27,45]. It is worth noting that these seemingly inconsistent results should not necessarily be viewed as a failure of SIF for monitoring plant water stress, as the changes in canopy structure may also directly reflect the plant water status. The inconsistency also highlights the need to consider/normalize canopy structure factors during drought. Otherwise, we may run the risk of wrongly assigning physiological causality to variance in TOC SIF due to changes in the canopy structure [28].
Equation (3) also contains other parameters (for example, KDF, Kmc, Kmc, Г* and ΦPmax). These additional parameters are often assumed to be constant in the literature but may actually vary with water stress. The actual KDF value is currently unknown. Gu et al. (2019) [31] set KDF to 19, while Liu et al. (2022) [46] assumed KDF to be 9. According to Equation (3), using these two KDF values will directly cause Anet to change by a factor of two. KDF remains unchanged at the same temperature, but there is currently no relevant research on KDF under water stress. Kmc, Kmc and Г* depend on the partial pressure of oxygen and temperature, which are related to the specificity factor of Rubisco [47]. Furthermore, ΦPmax decreases under water stress conditions. Considering these potential effects of water stress on the parameters involved in our theoretical equations, more future measurements are needed to quantitatively examine their relationships under water conditions. We performed a sensitivity analysis on the MLR model. The input variables are Ci, Kco, Γ*, ΦPmax, qL, and ChlFP_F, respectively. The sensitivity analysis results show that the main parameters affecting Anet variation are leaf physiological parameters, qL and ChlFP_F. These two parameters explain more than 80% of the Anet variation, while other leaf physiological parameters with greater influence such as Ci, Kco, and ΦPmax account for more than 15% of the total variation. However, the parameter Γ* plays a small role in explaining the Anet variation. Han et al. (2022) [47] studied the relationship between SIF and GPP through the MLR model and showed that the qL is sensitive to light intensity and can be expressed by the exponential equation of two parameters (aqL and bqL) between the qL and PAR. However, we found that in the uniform plant functional types, the parameters aqL and bqL vary greatly. Therefore, we use the NPQ/ΦP version of the Ja-ChlFP_F equation. The NPQ/ΦP version mixes photochemistry and non-photochemistry. Further, this mixing is superficial because information on non-photochemistry is canceled out in the product of (1 + NPQ) and ΦP/(1 − ΦP) as ΦP contains both photochemical and non-photochemical information. We need to consider both the energy-dependent and energy-independent components of NPQ. This is particularly important for this study because it focuses on water stress, which likely induces energy-independent NPQ.
The weak response of fluorescence to a water deficit suggests that SIF alone, at least its physiological component, is not able to track drought-induced changes in plant physiology. However, the development of the rMLR model allows for a mechanistic understanding of the drought impact on photosynthetic characteristics using SIF, soil moisture, and two measurable meteorological variables as the inputs. To apply the rMLR model at the canopy scale, that is, by using narrowband SIFtoc as an input variable for Equation (3), several more steps will need to be performed. First, the contribution of PSII fluorescence to SIFtoc (fPSII) should be determined. Although Bacour et al. [48] showed that fPSII can be also estimated from Tair and PAR, more research is needed to assess how fPSII varies with species and environmental conditions [44]. Second, the probability of a fluorescence photon escaping from a leaf level to canopy scale (fesc_L-C) must be quantified, in addition to fesc_P-L (Equation (8)). In the near-infrared (NIR) region, fesc_L-C can be estimated from directional reflectance (RNIR) [49]. Note that both SIFtoc and RNIR can be concurrently obtained from measurements of irradiance/radiance [50]. Thus, the estimation of fesc_L-C requires no additional observations. Third, a full-band SIF emission should be reconstructed from narrowband SIF. The full SIF spectrum at TOC can be approximated by a linear combination of basis spectra [46,51,52]. The Soil-Canopy Observations of Photosynthesis and Energy Balance (SCOPE) [53] model is typically used to generate a dataset that is representative of the majority of actual scenes. Principal Component Analysis (PCA) or Singular Value Decomposition (SVD) techniques are then applied to extract the basis spectra from this simulated dataset. However, the SCOPE model is designed for homogeneous vegetation canopies, such as crops, and its performance may deteriorate for a heterogeneous, structurally complex canopy [54]. Finally, a large-scale, and near-real-time root-zone soil moisture (RZSM) dataset is needed to estimate βS and βB. Since satellite soil moisture satellite observations are sensitive to surface soil moisture, typically within the first few centimeters, current RZSM datasets are mostly produced by assimilating observations into model simulations [55,56].

4. Materials and Methods

4.1. Leaf-Scale Concurrent Instrumentation

We developed a leaf-scale concurrent measurement system by integrating a portable gas-exchange system (LI-COR Biosciences, Lincoln, NE, USA), two HR2000+ spectrometers and two QE Pro spectrometers (Ocean Optics, Dunedin, FL, USA), a PAM system (Dual-PAM-100, Heinz Walz GmbH, Effeltrich, Germany), a short-pass filter, an external LED light source and fiber optics (connecting the PAM and spectrometers to the leaf chamber) (Figure 11). With this measurement system, we were able to simultaneously measure gas-exchange, passive and active ChlF, reflectance and transmittance for plants under a variety of controlled environmental conditions [46,57]. The main components and modifications are discussed below.

4.1.1. Gas-Exchange System

We used the LI-6800 gas-exchange system to control the environmental conditions of a modified leaf chamber. The original film of the LI-6800 transparent leaf chamber was replaced with high-transparency plexiglass. An external LED light source (see below) was fixed to the high-transparency plexiglass via a cylindrical plastic tube. The optical fibers of the PAM fluorometer and spectrometer were inserted into the leaf chamber via two fiber adapter bulkheads added to the plastic tube. These optical fibers were fixed to the diagonal of the fiber adapter bulkhead. In order to measure downward chlorophyll fluorescence spectral radiant energy fluxes, a metal plate covering the base of the leaf chamber was first added, and a fiber adapter bulkhead for inserting optical fibers was then fixed in the middle of this plate. All the fiber adapter bulkheads were sealed with clay to ensure the leaf chamber remained air-tight. To reduce light scattering in the leaf chamber, the inside of the light-source plastic tube and the upper surface of the metal plate connected to the base of the leaf chamber were painted with black acrylic paint (Black 2.0, Stuart Semple, UK) as a light trap (Figure 11b). The modified gas-exchange system blocks all light, except for incoming light with wavelengths longer than 625 nm, using a short-pass filter.

4.1.2. Gas-Exchange System

A Dual-PAM-100 instrument (Heinz Walz GmbH, Effeltrich, Germany) was used to measure fluorescence parameters. The Dual-PAM-100 featured a single red (625 nm) power LED for excitation of chlorophyll fluorescence. The saturating pulse was also driven by the red LED, which emitted a consistent 10,000 μmol m−2 s−1 for 0.8 s. The Dual-PAM-100 induced the maximum fluorescence quantum yield using the saturation pulse under dark-adapted conditions. The PAM was connected to the leaf chamber via a 1 m long single fiber of 0.8 cm diameter routed through the entrance hole (Figure 11a). The angle between the PAM fiber tip inserted into the entrance hole and the plane of the leaf chamber was kept at about 90° and the vertical distance between the fiber optic head and the surface of the leaf sample was maintained at 1 cm to avoid blocking the field of view of the PAM instrument.

4.1.3. Spectrometers

Using two HR2000+ and two QE Pro (Ocean Optics, Dunedin, FL, USA) high-sensitivity spectrometers, we simultaneously measured reflection, transmission, and upward and downward chlorophyll fluorescence spectral radiant energy fluxes of the leaf (Figure 11c). The leaf chamber was linked to the four spectrometers by means of two bifurcated optical fibers. One of these fibers was inserted into the top fiber adapter bulkhead above the leaf chamber to connect to an HR and a QE Pro, while the other was inserted into the bottom of the leaf chamber and connected to the other HR and QE Pro. Both bifurcated fibers had a diameter of 1000 um to ensure they could collect enough light. The vertical distance between the tip of the top fiber and the sample surface of the leaf was kept at 10 mm to allow for measurement of the upward chlorophyll fluorescence. The lower fiber optic head was inserted vertically into the base of the leaf chamber to measure the downward chlorophyll fluorescence, with the distance from the back of the leaf being 10 mm. The HR2000+ spectrometers covered the 296–1203 nm range at an optical resolution of 5.3159 nm with a spectral sampling of 0.4430 nm. The detector of the QE Pro spectrometer covered the range 634–863 nm at an optical resolution of 5.2656 nm with a spectral sampling of 0.2194 nm. The absolute calibration of the spectrometers and all light paths were carried out by using a separate reference QE Pro spectrometer (Ocean Optics, Dunedin, FL, USA) calibrated using a NIST traceable integration sphere. Dark current and non-linear spectrometer calibrations were completed before each measurement.

4.1.4. Light Source

The external actinic light source was a white LED light source (S5000, Nanjing Hecho Technology Co., Nanjing, China) with a ring-shaped fiber which provided a homogeneous light distribution across the leaf chamber. The actinic light source was capable of delivering 0–3000 μmol m−2 s−1 with a 400–700 nm wavelength range. To attenuate this LED light source, a customized 625 nm short-pass, fused silica filter with a diameter of 12.5 mm was placed between it and the O-ring fiber optic (Figure 11). After short-pass filtering, the rejection wavelength range of the external light source was, in fact, 639–925 nm, due to insufficient accuracy. In the 350–612 nm range, the transmittance exceeded 91% and the optical density was 4 (Edmunds Optics, Barrington, NJ, USA). A tin-foil lampshade outside the cylindrical plastic tube was used to block outside light. The light intensity was varied in the range of 0–2500 μmol m−2 s−1 by adjusting the pulse-width modulation controller in the light source.

4.2. Experiment Design

The experiment was conducted between 15 October 2020 and 30 April 2021 in the Northwest A&F University, Yangling, China. Winter wheat (Triticum aestivum L. cultivar Xi Nong 979) was sown in plastic pots (27 cm height × 21 cm diameter) filled with 9 kg of sieved, air-dried, loess topsoil and 1.95 g of urea. The field capacity of the soil was 24% and the wilting point was 9%. The plants were grown under a rainproof shed until the turning-green stage, and then moved to a climate chamber where CO2 concentration, air temperature and relative humidity were controlled at 400 µmol mol−1, 12 °C and 60%. The light intensity above the canopy during the experiment was kept at 600 μmol m−2 s−1 for 12 h per day, from 8 a.m. to 8 p.m. Beginning in late March, eight representative wheat plants in the heading stage were chosen for the experiment.
Throughout the 28-day experiment, the gravimetric soil water content (θSWC, %) was continuously monitored using the weighing method [58,59]. The wheat plants were subjected to one of two treatments: non-water stress (NS) and water stress (WS). Four wheat plants in each treatment were randomly selected as biological replicates. In the NS treatment, the θSWC was maintained at 19.1% (well-watered plants, assuming 80% of field capacity) [60,61] throughout the experiment, but, in the WS treatment, to simulate intensifying drought, θSWC was gradually decreased to a final value of 6.3%. This steady decline in θSWC ensured that plants under the WS treatment were subjected to progressive drought stress [62].
Measurements were made on attached leaves of the wheat plants under the two water treatments. The wheat leaf had to be placed in the 3 cm× 3 cm clear chamber along the diagonal of the square and positioned in the chamber with its adaxial surface facing the LED light source. Since the leaves did not fill the leaf chamber, the area of each leaf had to be accurately measured. After 1 h of dark adaptation, a saturation flash from the PAM fluorometer was used to determine minimal fluorescence (Fo) and maximal fluorescence in the dark (Fm) of dark-adapted leaves. Next, the light–response curves and CO2 response curves of gas exchange and fluorescence were measured. The CO2 flow rate, air relative humidity and leaf temperature were kept constant at 500 μmol s−1, 50% and 12 °C, respectively. To obtain light–response curves, measurements were conducted at a CO2 concentration of 400 µmol mol−1. Light–response measurements were made with photochemically active radiation (PAR) light intensities of 0, 40, 90, 180, 350, 700, 1300, 1700, and 2100 µmol m−2 s−1. Next, we measured the CO2 response curve under saturated light intensity (1500 µmol m−2 s−1) using values of the CO2 concentration gradient of 30, 50, 100, 200, 300, 400, 600, 900, 1200, 1500 μmol mol−1.
Throughout the course of each light–response curve and CO2 response curve determination, spectral measurements (reflected radiance, transmitted radiance, forward and backward fluorescence spectral radiant energy flux) were continually recorded at 1 s intervals. Steady-state fluorescence emission (Ft), induced by the measuring beam of the PAM fluorometer, was also included. Maximal fluorescence emission in the light-adapted state (Fm’) from the PAM was recorded under each light intensity and CO2 concentration. From these measurements (Fo, Ft, Fm, Fm’), we also obtained NPQ, a fraction of open PSII reaction centers (qL) [63], and the actual rate of electron transport (Ja_PAM, μmol m−2 s−1). The details for estimating them are provided in Note S1. Net CO2 assimilation rate (Anet, μmol m−2 s−1) and stomatal conductance to water vapor (GS, mol m−2 s−1) provided by the gas-exchange system were automatically stored every 5 s. To ensure that the gas-exchange conditions were stable, the measurements for light curves and CO2 curves were made after waiting at least 5 min, and up to 20 min, between each light intensity or CO2 concentration change. The filtered incident radiation of the LED light source at each light intensity were measured with a standard reflectance panel (Spectralon; Labsphere, North Sutton, NH, USA). The above measurements were taken at day 0 (before the water stress treatment started, WS0) and at days 2, 4, 5, 6, 7, 8, 9, 10, 12, 14, 16, 18, 20, 22, 24, 26, 28 after withholding water (WS2, WS4, etc., respectively). At WS0, the unstressed maximum carboxylation capacity of Rubisco (Vcmax,0, μmol m−2 s−1) and the unstressed maximum electron transport rate (Jmax,0, μmol m−2 s−1) were estimated by fitting the FvCB model [64] to the CO2 response curves. According to the photosynthetic light–response curve of winter wheat at WS0, Anet increased steeply with PAR when PAR ≤ 350 μmol m−2 s−1, and increased slowly with increasing PAR when PAR ≥ 1000 μmol m−2 s−1 (see below). Thus, a PAR level between 350 and 1000 μmol m−2 s−1 was assumed to be the intermediate light condition, and PAR levels lower than 350 or higher than 1000 μmol m−2 s−1 were taken to be low and high light conditions, respectively.

4.3. The Reformulated MLR (rMLR) Model

The MLR model [31] shows that Ja can be mechanistically estimated from qL, ФPmax, and the chlorophyll fluorescence flux density emitted from the photosystem II (PSII) across the full ChlF emission spectrum (ChlFP_F, μmol m−2 s−1). At the photosystem level,
J a = q L × Φ Pmax × ( 1 + K DF ) × ChlF P _ F ( 1 Φ Pmax ) ,
where KDF is the ratio between the rate constants for constitutive heat loss (KD) and fluorescence (KF) and is assumed to be 9 [46]. However, qL is much less studied than the other PAM parameters [31]. A previous study [46] shows that the role of qL and ФPmax in the original MLR model can be replaced by ФP and NPQ:
J a = Φ P × ( 1 + NPQ ) × ( 1 + K DF ) × ChlF P _ F ( 1 Φ P ) ,
The practical advantage of Equation (2) is that both ФP and NPQ (dimensionless) can be estimated from air temperature (Tair, °C) and PAR (see the section on the estimation of ФP and NPQ). Note that Ja in Equations (1) and (2) is already balanced by carboxylation and photorespiration [31,47]. To differentiate from Ja_PAM, in this study, Ja was calculated from ChlF emission using Equation (2).
One can obtain Anet for C3 and C4 species:
A net = A g R d = { C c Γ * 4 C c + 8 Γ * × Φ P × ( 1 + NPQ ) × ( 1 + K DF ) × ChlF P _ F ( 1 Φ P ) R d   C 3   1 ζ 3 × Φ P × ( 1 + NPQ ) × ( 1 + K DF ) × ChlF P _ F ( 1 Φ P ) R d     C 4   ,
where Ag represents gross photosynthesis (μmol m−2 s−1); Rd is the daytime respiration (μmol m−2 s−1); Cc is the chloroplastic CO2 partial pressure (µmol mol−1); Γ* is the chloroplastic compensation point of CO2 µmol mol−1 [31,65]; ζ is the fraction of total electron transport of mesophyll and bundle sheath allocated to mesophyll, assumed to be 0.4 [66]. We refer to Equation (3) as the rMLR model. The procedure for quantifying Anet from the observed SIF is illustrated in Figure 12.

4.4. Correction for the PSI Fluorescence

The rMLR model is only valid for fluorescence emissions from PSII [31]; the contribution of the PSI fluorescence should be excluded from all actively and passively induced ChlF related to Equation (3). Using the far-red method, Pfündel et al. [67] showed that the PSI fluorescence yield (F1) detected by the PAM fluorometer for C3 species can be estimated as:
F 1 = 0.24 × F o ,
By subtracting F1 from the fluorescence yields (Fo, Fm, Fm′, Ft) directly measured by the PAM fluorometer, we were able to correct them for PSI fluorescence. Accordingly, all other yields/parameters derived from these four yields, ФP, ФF, ФN, ФD, and NPQ, were also corrected for PSI fluorescence. See the details in Note S1. Hereafter, all of the active fluorescence parameters in this study only contain the PSII contribution unless otherwise specified.
The leaf-scale concurrent measurement system (Figure 11) provides the passive ChlF spectrum in the range 640 to 850 nm at the leaf scale (ChlFL(λ), mW m−2 nm−1 sr−1):
ChlF L ( λ ) = ChlF L _ U ( λ ) + ChlF L _ D ( λ ) ,
where ChlFL_U(λ) and ChlFL_D(λ) represent the fluorescence radiance emitted from adaxial and abaxial leaf surfaces (μW cm−2 nm−1 sr−1), respectively, and λ is the wavelength (nm).
Due to the strong linear relationships between spectral fluorescence yields and the original PAM Ft yields [68,69], deriving the ratio of Ft yields before and after correction for PSI fluorescence can also allow for an approximate separation of the PSI and PSII spectral fluorescence yields. Further, Pfündel [70] showed that the ratio of F1 to Fo was 14% and 45% in the spectral ranges below 700 nm (SW) and above 700 nm (LW), respectively. Considering the wavelength-dependent relationships among spectral and PAM fluorescence yields [69], the contribution of PSII (ChlFL_PSII(λ)) to measurements of ChlFL(λ) can be estimated as:
ChlF L _ PSII ( λ ) = { ChlF L ( λ ) × F t F 1 _ SW F t λ 700 nm ChlF L ( λ ) × F t F 1 _ LW F t λ > 700 nm ,
where Ft is the steady-state fluorescence yield, and F1_SW (F1_SW = 0.14 × Fo) and F1_LW (F1_LW = 0.45 × Fo) represent the PSI contribution at SW wavelengths and LW wavelengths, respectively. Note that Ft and Fo used in Equation (6) were directly measured by the PAM fluorometer and thus contain contributions from both PSI and PSII.
To apply the rMLR model, ChlFL_PSII(λ) must be further downscaled to the photosystem level (ChlFP(λ)) by accounting for the probability that a fluorescence photon escapes from the PSII light reactions inside the leaves to the surface of the leaf (fesc_P-L):
ChlF P ( λ ) = ( ChlF L _ PSII ( λ ) ) / f e s c _ P L ( λ ) ,
fesc_P-L is approximately equal to the sum of leaf reflectance (R) and transmittance (T) [71]:
f esc _ P - L R ( λ ) + T ( λ )   ,
Note that ChlFP(λ) in Equation (7) has units of mW m−2 nm−1 sr−1. To obtain ChlFP_F as required by Equation (3), we need to integrate ChlFP(λ) between 640 and 850 nm and perform a unit conversion:
ChlF P _ F = λ = 640 850 [ ChlF P ( λ ) ] × λ × 10 6 h × c × N A × 10 3 × 10 9   ,
where h is the Planck constant (6.62607015 × 10−34 J·s), c is the light speed (3 × 108 m s−1), NA is the Avogadro constant (6.02 × 1023 mol−1), 106 is used to convert moles (mol) to micromoles (μmol) in NA, 103 is used to convert milliwatts (mW) to Watts (W), and 109 is used to convert nanometers (nm) to meters (m) in λ. For the application of the rMLR model at the canopy scale or beyond, see the see the Section 3.

4.5. Estimation of ФP and NPQ

In this study, KD and KF are assumed to be 0.9 and 0.1, respectively [46]. Note that NPQ should be equal to the rate coefficients of energy-dependent heat dissipation (KN) because NPQ = KN/(KF + KD), and KF + KD = 1. KN can be estimated as [48]:
K N = a × χ c × 1 + b b + χ c × exp ( d × T air + e ) PAR f ,
where a, b, c, d, e and f are fitting parameters. Following Bacour, Maignan, MacBean, Porcar-Castell, Flexas, Frankenberg, Peylin, Chevallier, Vuichard and Bastrikov [48], b, c, d, e and f are assumed to have values of 5.74, 2.167, −0.014, −0.00437 and 0.00576, respectively. Potentially, we can use an exponential equation with three parameters to represent the relationship between the parameter a and θSWC:
a = g × exp ( ( h ) × θ swc ) + j ,
where g, h, and j are empirical parameters. The measurements are randomly divided into two groups, with 50% of the data used for training, and the remaining 50% for evaluating the performance of the predictions. The Matlab function ‘lsqnonlin’ was used to obtain the values of the parameters by minimizing a cost function on the training dataset: C = (M − S)2, where M is the NPQ measured with the PAM, and S is the modelled NPQ. A Trust Region algorithm was used to update the values of the parameters after each iteration step, with the iteration terminating when the improvement in the cost function was less than 10−3. Table S1 presents the initial values, boundaries and constraints of the parameters.
χ is defined as [32,46]:
χ = 1 Φ P Φ Pmax ,
where ФPmax is assumed to be 0.8 and is similar among healthy plants; ФP is estimated as [32]:
Φ P = min ( A C , A J ) α grn × β PSII × PAR × 4 C c + 8 Γ * C c Γ * ,
where AC and AJ represent Rubisco-limited and RuBP-limited gross CO2 assimilations (μmol m−2 s−1), respectively. αgrn represents the absorption efficiency of PAR by green leaves, and the value is usually fixed at 0.84. βPSІІ is the fraction of absorbed energy allocated to PSII, and the value is set to 0.5 [72]. AC and AJ are given by [73]:
A C = V cmax × ( C c Γ * ) C c + K mC × ( 1 + O c / K mO ) ,
A J = J p 4 ( C c Γ * ) C c + 2 Γ *
where KmC is the Michaelis–Menten constants of Rubisco for CO2 270 µbar, [73]; KmO is the Michaelis–Menten constants of Rubisco for O2 16,500 µbar, [73]; Oc is the chloroplastic O2 partial pressure, assumed to equal to the oxygen partial pressure 230,000 µbar, [74]; Jp is the potential electron transport rate [47]:
J p = σ × PAR + J max ( σ × PAR + J max ) 2 4 × θ × σ × PAR × J max 2 × θ ,
where σ is the product of leaf light absorptance, a fraction of absorbed photons allocated to PSII and ФPmax. σ is set to 0.3 [72]. θ is an empirical curvature parameter, which is also modelled as a function of θSWC:
θ = k × exp ( ( l ) × θ swc ) + m ,
where k, l, and m are the fitting parameters. Again, 50% of the measurements were used to determine the parameter values by minimizing the squared difference between the measured and simulated values of ФP (Table S1).
Anet is limited by biochemical processes under water stress, such that a soil-moisture-dependent stress function (βB, see Equation (22) below) should be applied to regulate the parameters Jmax and Vcmax of the photosynthesis model [75]:
J max = β B × J max , 0 ,
V cmax = β B × V cmax , 0
where Jmax,0 and Vcmax,0 represent the unstressed values of Jmax and Vcmax, respectively, at the beginning of the experiment.

4.6. Estimation of Γ*, Rd, and Cc

Katul et al. [76] and Liu et al. [77] showed that Γ* can be estimated as a function of Tair:
Γ * = 36 . 9 + 1 . 18 × ( T air 25 ) + 0 . 036 × ( T air 25 ) 2 ,
Here, we used air temperature as measured in the LI-6800 leaf chamber. Rd is described as [78]:
R d = 0.015 × V cmax , 0 ,
Mesophyll conductance to CO2 was assumed to be infinite and thus Cc was considered to be equal to intercellular CO2 partial pressure Ci, [53]. Ci is estimated as [79]:
C i = C a A net G c ,
where Ca is the ambient air CO2 partial pressure (µmol mol−1), and Gc is the stomatal conductance for CO2 (mol m−2 s−1). Anet is the minimum of AC and AJ (Equation (14)). Because Ci, Anet, and Gc are coupled to each other, the estimation of Anet and Gc has to be resolved iteratively over Ci given an initial value, which is Ci = 0.7 × Ca for C3 winter wheat [80]. The iterative loop stops when the difference in Ci between two successive iterations is less than 0.1 µmol mol−1. The biochemical model of photosynthesis proposed by Farquhar, von Caemmerer and Berry [64] was used to estimate Anet as the minimum of the Rubisco-limited CO2 assimilation rate and the electron-transport-limited CO2 assimilation rate.
Gc is estimated using a modified Ball–Woodrow–Berry model BWB [81]:
G s = G 0 + a × β S × A net C s × ( 1 + VPD / D 0 ) ,
G c = 0 . 64 × G s
where G0 is the residual conductance (mol m−2 s−1), assumed to be 0.01 [81]; Cs is the CO2 concentration at the leaf surface (µmol mol−1), assumed to be the product of a/(a − 1) and Ci; VPD is the vapor pressure deficit (kPa); D0 (kPa) is an empirical parameter related to stomatal sensitivity to VPD, assumed to be 1.5 [81]; a is a parameter related to Ci, assumed to be 11.0 [81]; 0.64 is a factor used to convert the molecular diffusivity of water vapor to CO2 [79]; βS is the normalized soil-moisture-dependent stress function which accounts for the reduction in GS under water stress. βS and βB (Equation (17)) can be defined as:
β i = {   0 θ SWC < θ WP [ θ SWC θ WP θ FC θ WP ] q j θ WP θ SWC θ FC ,   1 θ SWC > θ FC
where βi ranges between 1 (for plants not suffering from drought) and 0 (transpiration is zero); the subscripts i = B and S are for biochemical and stomatal limitations, respectively [75]; θFC and θWP represent θSWC at the field capacity (24%) and wilting point (9%); the fitting parameter qj is a measure of the nonlinearity of the effects of water stress on the biochemical and stomatal limitations, and j takes the values B and S for i = B and S, respectively [75,82]. qB (Equation (17)) and qS in (Equation (21)) were determined by minimizing a cost function, C = (M − S)2, where M represents the measurement and S is the corresponding simulated value (Vcmax for qB, and Gs for qS).

5. Conclusions

Our results show that the response of fluorescence emissions to drought is smaller than those of either stomatal conductance or net photosynthetic carbon assimilation. At the canopy scale and beyond, however, structural dynamics dominate the spatial variation of canopy SIF in response to water stress, explaining the strong drought response of SIF retrieved from space. As drought becomes more severe, the shifts in energy allocation towards decreasing photochemistry and increasing fluorescence emission tend to push plants into the PQ phase, enhancing the nonlinearity in the overall relationship between photochemistry and fluorescence. We confirm that SIF alone has a limited ability to predict drought-induced declines in photosynthetic parameters. Alternatively, the rMLR model, using SIF as one important input variable, demonstrates a satisfactory performance in reproducing declines in stomatal conductance and net photosynthetic carbon assimilation. The rMLR model has good potential for applications at regional and global scales, and thus provides the basis for using SIF mechanistically to estimate GPP under the scenario of increasing intensity and the extent of droughts in the twenty-first century.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/plants12193365/s1; Notes S1: Calculation of ФP, ФF, ФN, ФD, NPQ, ФPmax and Ja_PAM; Figure S1: Variations in βS and βB for the water stress (WS, diamonds) treatments under 12 days of progressive drought stress; Figure S2: Average absorbed photosynthetically active radiation (APAR, μmol m−2 s−1) under changing light intensity over the duration of the drought; Figure S3: The responses of full-band chlorophyll fluorescence emission at the photosystem level (ChlFP_F, μmol m−2 s−1) to changing light intensity during the period between 14 and 28 days (WS14 to WS28) after withholding water; Table S1: List of retrieved parameters, their initial values, lower boundaries (LB), upper boundaries (UB) and constraints.

Author Contributions

Q.J. and X.L. planned and designed the research. Z.L. developed model assumptions and scenarios. Q.J., C.G., Y.W., J.Y. and J.W. performed experiments. Q.J. analyzed data. Q.J. and X.L. wrote the manuscript. Q.Y., F.Z. and X.L. edited the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The research was supported by the National Natural Science Foundation of China (Grant No. 41901293 to Z.L., and 42071328 to X.L.), the China Postdoctoral Science Foundation (Grant No. 2019M663828 to Z.L.), and the Chinese Universities Scientific Fund (24520212452021125 to X.L.).

Data Availability Statement

The data are available on request from the authors. The data supporting the findings of this study are available from the corresponding author, Xiaoliang Lu, upon request.

Acknowledgments

The authors would like to thank the anonymous reviewers for their valuable comments and suggestions. The authors gratefully thank Huanlin Gao for helping with the experiment.

Conflicts of Interest

The authors have no conflict of interest to declare.

Appendix A. Definitions, Symbols and Methods for the Nomenclature in This Paper

SymbolsDefinitionUnit
Jathe actual rate of electron transport calculated from ChlF emissionμmol m−2 s−1
qLfraction of open PSII reaction centers/
ФPmaxmaximum photochemical quantum yield of PSII/
KDthe rate constants for constitutive heat loss/
KFthe rate constants for fluorescence/
KDFthe ratio between (KD) and (KF)/
ChlFP_Fchlorophyll fluorescence flux density emitted from the photosystem II (PSII) across the full ChlF emission spectrumμmol m−2 s−1
ФPthe quantum yields of photochemical quenching in PSII/
NPQnon-photochemical quenching,/
PARphotochemically active radiationμmol m−2 s−1
Tairair temperature°C
Anetnet CO2 assimilation rateμmol m−2 s−1
Aggross photosynthesisμmol m−2 s−1
Rdthe daytime respirationμmol m−2 s−1
Ccthe chloroplastic CO2 partial pressureμmol m−2 s−1
Γ*the chloroplastic compensation point of CO2μmol m−2 s−1
ζthe fraction of total electron transport of mesophyll and bundle sheath allocated to mesophyll/
F1the PSI fluorescence yield by the PAM fluorometer for C3 species/
Fominimal fluorescence in the dark of dark-adapted leaves from the PAM/
Fmmaximal fluorescence in the dark of dark-adapted leaves from the PAM/
Fmmaximal fluorescence emission in the light-adapted state from the PAM/
Ftsteady-state fluorescence emission induced by the measuring beam of the PAM fluorometer/
ФFthe quantum yield of fluorescence emission/
ФNthe quantum yield of regulated heat dissipation/
ФDthe quantum yield of constitutive heat dissipation/
ChlFL(λ)the passive ChlF spectrum in the range 640 to 850 nm at the leaf scalemW m−2 nm−1 sr−1
ChlFL_U(λ)the fluorescence radiance emitted from adaxial leaf surfacesμW cm−2 nm−1 sr−1
ChlFL_Dthe fluorescence radiance emitted from abaxial leaf surfacesμW cm−2 nm−1 sr−1
ChlFL_PSII(λ)the contribution of PSII to measurements of ChlFL(λ)mW m−2 nm−1 sr−1
F1_SWthe PSI contribution at SW wavelengths, F1_SW = 0.14 × Fo/
F1_LWthe PSI contribution at LW wavelengths, F1_LW = 0.45 × Fo/
ChlFP(λ)downscaling the ChlFL_PSII(λ) to the photosystem levelmW m−2 nm−1 sr−1
fesc_P-Lthe probability that a fluorescence photon escapes from the PSII light reactions inside the leaves to the surface of the leaf/
Rleaf reflectance/
Tleaf transmittance/
KNthe rate coefficients of energy-dependent heat dissipation/
χthe relative light saturation/
θSWCgravimetric soil water content%
ACRubisco-limited gross CO2 assimilationμmol m−2 s−1
AJRuBP-limited gross CO2 assimilationμmol m−2 s−1
αgrnthe absorption efficiency of PAR by green leaves/
βPSІІfraction of absorbed energy allocated to PSII/
σthe product of leaf light absorptance, fraction of absorbed photons allocated to PSII and ФPmax/
Vcmaxthe maximum carboxylation capacity of Rubiscoμmol m−2 s−1
KmCMichaelis–Menten constants of Rubisco for CO2µbar
KmoMichaelis–Menten constants of Rubisco for O2µbar
Octhe chloroplastic O2 partial pressureµbar
Jpthe potential electron transport rateμmol m−2 s−1
Jmaxthe maximum electron transport rateμmol m−2 s−1
βBthe soil-moisture-dependent stress function which accounts for the reduction in Vcmax and Jmax under water stress/
Vcmax,0unstressed Vcmax at the beginning of the experimentμmol m−2 s−1
Jmax,0unstressed Jmax at the beginning of the experimentμmol m−2 s−1
Cithe intercellular CO2 concentrationµmol mol−1
Cathe ambient air CO2 partial pressureµmol mol−1
Gcthe stomatal conductance for CO2mol m−2 s−1
GSstomatal conductance to water vapormol m−2 s−1
G0the residual conductancemol mol−1
aa parameter related to Ci/
βSthe normalized soil-moisture-dependent stress function which accounts for the reduction in GS under water stress/
Csthe CO2 concentration at the leaf surfaceµmol mol−1
VPDthe vapor pressure deficitkPa
D0an empirical parameter related to stomatal sensitivity to VPDkPa
qBa measure of the nonlinearity of the effects of water stress on the biochemical/
qSa measure of the nonlinearity of the effects of water stress on the stomatal limitations/
θFCθSWC at field capacity%
θWPθSWC at wilting point%
Ja_PAMthe actual rate of electron transport from the PAMμmol m−2 s−1
Jmax,0unstressed Jmax at the beginning of the experimentμmol m−2 s−1

References

  1. Baker, N.R. Chlorophyll fluorescence: A probe of photosynthesis in vivo. Annu. Rev. Plant Biol. 2008, 59, 89–113. [Google Scholar] [CrossRef] [PubMed]
  2. Meroni, M.; Rossini, M.; Guanter, L.; Alonso, L.; Rascher, U.; Colombo, R.; Moreno, J. Remote sensing of solar-induced chlorophyll fluorescence: Review of methods and applications. Remote Sens. Environ. 2009, 113, 2037–2051. [Google Scholar] [CrossRef]
  3. 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]
  4. Magney, T.S.; Frankenberg, C.; Köhler, P.; North, G.; Davis, T.S.; Dold, C.; Dutta, D.; Fisher, J.B.; Grossmann, K.; Harrington, A. Disentangling changes in the spectral shape of chlorophyll fluorescence: Implications for remote sensing of photosynthesis. J. Geophys. Res. Biogeosci. 2019, 124, 1491–1507. [Google Scholar] [CrossRef]
  5. Schreiber, U.; Schliwa, U.; Bilger, W. Continuous recording of photochemical and non-photochemical chlorophyll fluorescence quenching with a new type of modulation fluorometer. Photosynth. Res. 1986, 10, 51–62. [Google Scholar] [CrossRef]
  6. Bilger, W.; Schreiber, U.; Bock, M. Determination of the quantum efficiency of photosystem II and of non-photochemical quenching of chlorophyll fluorescence in the field. Oecologia 1995, 102, 425–432. [Google Scholar] [CrossRef]
  7. Govindjee, G. Sixty-three years since Kautsky: Chlorophyll a fluorescence. Aust. J. Plant Physiol. 1995, 22, 131–160. [Google Scholar] [CrossRef]
  8. 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]
  9. Yang, X.; Tang, J.W.; Mustard, J.F.; Lee, J.E.; Rossini, M.; Joiner, J.; Munger, J.W.; Kornfeld, A.; Richardson, A.D. Solar-induced chlorophyll fluorescence that correlates with canopy photosynthesis on diurnal and seasonal scales in a temperate deciduous forest. Geophys. Res. Lett. 2015, 42, 2977–2987. [Google Scholar] [CrossRef]
  10. Köhler, P.; Frankenberg, C.; Magney, T.S.; Guanter, L.; Joiner, J.; Landgraf, J. Global retrievals of solar-induced chlorophyll fluorescence with TROPOMI: First results and intersensor comparison to OCO-2. Geophys. Res. Lett. 2018, 45, 10456–10463. [Google Scholar] [CrossRef]
  11. Yang, H.L.; Yang, X.; Zhang, Y.G.; Heskel, M.A.; Lu, X.L.; Munger, J.W.; Sun, S.; Tang, J.W. Chlorophyll fluorescence tracks seasonal variations of photosynthesis from leaf to canopy in a temperate forest. Glob. Chang. Biol. 2017, 23, 2874–2886. [Google Scholar] [CrossRef] [PubMed]
  12. He, L.Y.; Magney, T.; Dutta, D.; Yin, Y.; Köhler, P.; Grossmann, K.; Stutz, J.; Dold, C.; Hatfield, J.; Guan, K.Y. From the ground to space: Using solar-induced chlorophyll fluorescence to estimate crop productivity. Geophys. Res. Lett. 2020, 47, e2020GL087474. [Google Scholar] [CrossRef]
  13. 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. 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]
  14. Verma, M.; Schimel, D.; Evans, B.; Frankenberg, C.; Beringer, J.; Drewry, D.T.; Magney, T.; Marang, I.; Hutley, L.; Moore, C. Effect of environmental conditions on the relationship between solar-induced fluorescence and gross primary productivity at an OzFlux grassland site. J. Geophys. Res. Biogeosci. 2017, 122, 716–733. [Google Scholar] [CrossRef]
  15. Sun, Y.; Frankenberg, C.; Jung, M.; Joiner, J.; Guanter, L.; Köhler, P.; Magney, T. Overview of Solar-Induced chlorophyll Fluorescence (SIF) from the Orbiting Carbon Observatory-2: Retrieval, cross-mission comparison, and global monitoring for GPP. Remote Sens. Environ. 2018, 209, 808–823. [Google Scholar] [CrossRef]
  16. Marrs, J.; Reblin, J.; Logan, B.; Allen, D.; Reinmann, A.; Bombard, D.; Tabachnik, D.; Hutyra, L. Solar-induced fluorescence does not track photosynthetic carbon assimilation following induced stomatal closure. Geophys. Res. Lett. 2020, 47, e2020GL087956. [Google Scholar] [CrossRef]
  17. IPCC. Climate Change and Land: An IPCC Special Report on Climate Change, Desertification, Land Degradation, Sustainable Land Management, Food Security, and Greenhouse Gas Fluxes in Terrestrial Ecosystems; IPCC: Geneva, Switzerland, 2019. [Google Scholar]
  18. Hanjra, M.A.; Qureshi, M.E. Global water crisis and future food security in an era of climate change. Food Policy 2010, 35, 365–377. [Google Scholar] [CrossRef]
  19. Kim, W.; Iizumi, T.; Nishimori, M. Global patterns of crop production losses associated with droughts from 1983 to 2009. J. Appl. Meteorol. Climatol. 2019, 58, 1233–1244. [Google Scholar] [CrossRef]
  20. Leng, G.Y.; Hall, J. Crop yield sensitivity of global major agricultural countries to droughts and the projected changes in the future. Sci. Total Environ. 2019, 654, 811–821. [Google Scholar] [CrossRef]
  21. Ayaz, A.; Huang, H.; Zheng, M.; Zaman, W.; Li, D.; Saqib, S.; Zhao, H.; Lü, S. Molecular cloning and functional analysis of GmLACS2-3 reveals its involvement in cutin and suberin biosynthesis along with abiotic stress tolerance. Int. J. Mol. Sci. 2021, 22, 9175. [Google Scholar] [CrossRef] [PubMed]
  22. FAOSTAT. Food and Agriculture Organization of the United Nations—Crop Statistics on Lupin Production; FAOSTAT: Rome, Italy, 2018. [Google Scholar]
  23. Lu, C.M.; Zhang, J.H. Effects of water stress on photosystem II photochemistry and its thermostability in wheat plants. J. Exp. Bot. 1999, 50, 1199–1206. [Google Scholar] [CrossRef]
  24. Hlavinka, P.; Trnka, M.; Semerádová, D.; Dubrovský, M.; Žalud, Z.; Možný, M. Effect of drought on yield variability of key crops in Czech Republic. Agric. For. Meteorol. 2009, 149, 431–442. [Google Scholar] [CrossRef]
  25. Ouyang, W.J.; Struik, P.C.; Yin, X.Y.; Yang, J.C. Stomatal conductance, mesophyll conductance, and transpiration efficiency in relation to leaf anatomy in rice and wheat genotypes under drought. J. Exp. Bot. 2017, 68, 5191–5205. [Google Scholar] [CrossRef]
  26. Sun, Y.; Fu, R.; Dickinson, R.; Joiner, J.; Frankenberg, C.; Gu, L.H.; Xia, Y.L.; Fernando, N. Drought onset mechanisms revealed by satellite solar-induced chlorophyll fluorescence: Insights from two contrasting extreme events. J. Geophys. Res. Biogeosci. 2015, 120, 2427–2440. [Google Scholar] [CrossRef]
  27. Yoshida, Y.; Joiner, J.; Tucker, C.; Berry, J.; Lee, J.-E.; Walker, G.; Reichle, R.; Koster, R.; Lyapustin, A.; Wang, Y. The 2010 Russian drought impact on satellite measurements of solar-induced chlorophyll fluorescence: Insights from modeling and comparisons with parameters derived from satellite reflectances. Remote Sens. Environ. 2015, 166, 163–177. [Google Scholar] [CrossRef]
  28. Xu, S.; Atherton, J.; Riikonen, A.; Zhang, C.; Oivukkamäki, J.; MacArthur, A.; Honkavaara, E.; Hakala, T.; Koivumäki, N.; Liu, Z.G. Structural and photosynthetic dynamics mediate the response of SIF to water stress in a potato crop. Remote Sens. Environ. 2021, 263, 112555. [Google Scholar] [CrossRef]
  29. Helm, L.T.; Shi, H.Y.; Lerdau, M.T.; Yang, X. Solar-induced chlorophyll fluorescence and short-term photosynthetic response to drought. Ecol. Appl. 2020, 30, e02101. [Google Scholar] [CrossRef]
  30. Magney, T.S.; Barnes, M.L.; Yang, X. On the covariation of chlorophyll fluorescence and photosynthesis across scales. Geophys. Res. Lett. 2020, 47, e2020GL091098. [Google Scholar] [CrossRef]
  31. Gu, L.H.; Han, J.M.; 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]
  32. van der Tol, C.; Berry, J.; Campbell, P.; 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]
  33. 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]
  34. Maguire, A.J.; Eitel, J.U.; Griffin, K.L.; Magney, T.S.; Long, R.A.; Vierling, L.A.; Schmiege, S.C.; Jennewein, J.S.; Weygint, W.A.; Boelman, N.T. On the functional relationship between fluorescence and photochemical yields in complex evergreen needleleaf canopies. Geophys. Res. Lett. 2020, 47, e2020GL087858. [Google Scholar] [CrossRef]
  35. Chen, X.J.; Mo, X.G.; Hu, S.; Liu, S.X. Relationship between fluorescence yield and photochemical yield under water stress and intermediate light conditions. J. Exp. Bot. 2019, 70, 301–313. [Google Scholar] [CrossRef] [PubMed]
  36. Faraloni, C.; Cutino, I.; Petruccelli, R.; Leva, A.R.; Lazzeri, S.; Torzillo, G. Chlorophyll fluorescence technique as a rapid tool for in vitro screening of olive cultivars (Olea europaea L.) tolerant to drought stress. Environ. Exp. Bot. 2011, 73, 49–56. [Google Scholar] [CrossRef]
  37. Jiang, Y.; Ye, J.; Niinemets, Ü. Dose-dependent methyl jasmonate effects on photosynthetic traits and volatile emissions: Biphasic kinetics and stomatal regulation. Plant Signal. Behav. 2021, 16, 1917169. [Google Scholar] [CrossRef]
  38. Jiang, Y.; Ye, J.; Rasulov, B.; Niinemets, Ü. Role of stomatal conductance in modifying the dose response of stress-volatile emissions in methyl jasmonate treated leaves of cucumber (Cucumis sativa). Int. J. Mol. Sci. 2020, 21, 1018. [Google Scholar] [CrossRef]
  39. Yamashita, F.; Rodrigues, A.L.; Rodrigues, T.M.; Palermo, F.H.; Baluška, F.; de Almeida, L.F.R. Potential plant–plant communication induced by infochemical methyl jasmonate in sorghum (Sorghum bicolor). Plants 2021, 10, 485. [Google Scholar] [CrossRef] [PubMed]
  40. Muller, P.; Li, X.P.; Niyogi, K.K. Non-photochemical quenching. A response to excess light energy. Plant Physiol. 2001, 125, 1558–1566. [Google Scholar] [CrossRef] [PubMed]
  41. Flexas, J.; Medrano, H. Drought-inhibition of photosynthesis in C3 plants: Stomatal and non-stomatal limitations revisited. Ann. Bot. 2002, 89, 183–189. [Google Scholar] [CrossRef]
  42. Adams, W.W.; Zarter, C.R.; Ebbert, V.; Demmig-Adams, B. Photoprotective strategies of overwintering evergreens. Bioscience 2004, 54, 41–49. [Google Scholar] [CrossRef]
  43. Verhoeven, A. Sustained energy dissipation in winter evergreens. New Phytol. 2013, 201, 57–65. [Google Scholar] [CrossRef]
  44. Porcar-Castell, A.; Malenovský, Z.; Magney, T.; Van Wittenberghe, S.; Fernández-Marín, B.; Maignan, F.; Zhang, Y.G.; Maseyk, K.; Atherton, J.; Albert, L.P. Chlorophyll a fluorescence illuminates a path connecting plant molecular biology to earth-system science. Nat. Plants 2021, 7, 998–1009. [Google Scholar] [CrossRef] [PubMed]
  45. Sloat, L.L.; Lin, M.; Butler, E.E.; Johnson, D.; Holbrook, N.M.; Huybers, P.J.; Lee, J.-E.; Mueller, N.D. Evaluating the benefits of chlorophyll fluorescence for in-season crop productivity forecasting. Remote Sens. Environ. 2021, 260, 112478. [Google Scholar] [CrossRef]
  46. Liu, Z.Q.; Zhao, F.; Liu, X.J.; Yu, Q.; Wang, Y.F.; Peng, X.B.; Cai, H.J.; Lu, X.L. Direct estimation of photosynthetic CO2 assimilation from solar-induced chlorophyll fluorescence (SIF). Remote Sens. Environ. 2022, 271, 112893. [Google Scholar] [CrossRef]
  47. Han, J.M.; Chang, C.Y.Y.; Gu, L.H.; Zhang, Y.J.; Meeker, E.W.; Magney, T.S.; Walker, A.P.; Wen, J.M.; Kira, O.; McNaull, S.; et al. The physiological basis for estimating photosynthesis from chlorophyll a fluorescence. New Phytol. 2022, 234, 120–1219. [Google Scholar] [CrossRef] [PubMed]
  48. Bacour, C.; Maignan, F.; MacBean, N.; Porcar-Castell, A.; Flexas, J.; Frankenberg, C.; Peylin, P.; Chevallier, F.; Vuichard, N.; Bastrikov, V. Improving estimates of gross primary productivity by assimilating solar-induced fluorescence satellite retrievals in a terrestrial biosphere model using a process-based SIF model. J. Geophys. Res. Biogeosci. 2019, 124, 3281–3306. [Google Scholar] [CrossRef]
  49. Zeng, Y.L.; Badgley, G.; Dechant, B.; Ryu, Y.; Chen, M.; 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]
  50. Lu, X.L.; Liu, Z.Q.; Zhao, F.; Tang, J.W. Comparison of total emitted solar-induced chlorophyll fluorescence (SIF) and top-of-canopy (TOC) SIF in estimating photosynthesis. Remote Sens. Environ. 2020, 251, 112083. [Google Scholar] [CrossRef]
  51. Zhao, F.; Guo, Y.Q.; Verhoef, W.; Gu, X.F.; Liu, L.Y.; Yang, G.J. A method to reconstruct the solar-induced canopy fluorescence spectrum from hyperspectral measurements. Remote Sens. 2014, 6, 10171–10192. [Google Scholar] [CrossRef]
  52. Zhao, F.; Li, R.; Verhoef, W.; Cogliati, S.; Liu, X.J.; Huang, Y.B.; Guo, Y.Q.; Huang, J.X. Reconstruction of the full spectrum of solar-induced chlorophyll fluorescence: Intercomparison study for a novel method. Remote Sens. Environ. 2018, 219, 233–246. [Google Scholar] [CrossRef]
  53. van der Tol, C.; Verhoef, W.; Rosema, A. A model for chlorophyll fluorescence and photosynthesis at leaf scale. Agric. For. Meteorol. 2009, 149, 96–105. [Google Scholar] [CrossRef]
  54. Mohammed, G.H.; Colombo, R.; Middleton, E.M.; Rascher, U.; van der Tol, C.; Nedbal, L.; Goulas, Y.; Pérez-Priego, O.; Damm, A.; Meroni, M. Remote sensing of solar-induced chlorophyll fluorescence (SIF) in vegetation: 50 years of progress. Remote Sens. Environ. 2019, 231, 111177. [Google Scholar] [CrossRef]
  55. Rodell, M.; Houser, P.R.; Jambor, U.; Gottschalck, J.; Mitchell, K.; Meng, C.-J.; Arsenault, K.; Cosgrove, B.; Radakovich, J.; Bosilovich, M. The global land data assimilation system. Bull. Am. Meteorol. Soc. 2004, 85, 381–394. [Google Scholar] [CrossRef]
  56. Reichle, R.H.; Liu, Q.; Koster, R.D.; Crow, W.T.; De Lannoy, G.J.; Kimball, J.S.; Ardizzone, J.V.; Bosch, D.; Colliander, A.; Cosh, M. Version 4 of the SMAP Level-4 soil moisture algorithm and data product. J. Adv. Model. Earth Syst. 2019, 11, 3106–3130. [Google Scholar] [CrossRef]
  57. Meeker, E.W.; Magney, T.S.; Bambach, N.; Momayyezi, M.; McElrone, A.J. Modification of a gas exchange system to measure active and passive chlorophyll fluorescence simultaneously under field conditions. AoB Plants 2021, 13, plaa066. [Google Scholar] [CrossRef] [PubMed]
  58. Chen, Y.G.; Chen, F.Q.; Liu, L.; Zhu, S.B. Physiological responses of Leucaena leucocephala seedlings to drought stress. Procedia Eng. 2012, 28, 110–116. [Google Scholar]
  59. Wei, Y.Q.; Jin, J.L.; Jiang, S.M.; Ning, S.W.; Liu, L. Quantitative response of soybean development and yield to drought stress during different growth stages in the Huaibei Plain, China. Agronomy 2018, 8, 97. [Google Scholar] [CrossRef]
  60. Li, F.S.; Liang, J.H.; Kang, S.Z.; Zhang, J.H. Benefits of alternate partial root-zone irrigation on growth, water and nitrogen use efficiencies modified by fertilization and soil water status in maize. Plant Soil 2007, 295, 279–291. [Google Scholar] [CrossRef]
  61. Arzanesh, M.H.; Alikhani, H.A.; Khavazi, K.; Rahimian, H.A.; Miransari, M. Wheat (Triticum aestivum L.) growth enhancement by Azospirillum sp. under drought stress. World J. Microbiol. Biotechnol. 2011, 27, 197–205. [Google Scholar] [CrossRef]
  62. Dossa, K.; Diouf, D.; Cissé, N. Genome-wide investigation of Hsf genes in sesame reveals their segmental duplication expansion and their active role in drought stress response. Front. Plant Sci. 2016, 7, 1522. [Google Scholar] [CrossRef]
  63. Kramer, D.M.; Johnson, G.; Kiirats, O.; Edwards, G.E. New fluorescence parameters for the determination of QA redox state and excitation energy fluxes. Photosynth. Res. 2004, 79, 209–218. [Google Scholar] [CrossRef]
  64. Farquhar, G.D.; von Caemmerer, S.; Berry, J.A. A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta 1980, 149, 78–90. [Google Scholar] [CrossRef] [PubMed]
  65. Long, S.P.; Bernacchi, C. Gas exchange measurements, what can they tell us about the underlying limitations to photosynthesis? Procedures and sources of error. J. Exp. Bot. 2003, 54, 2393–2401. [Google Scholar] [CrossRef] [PubMed]
  66. von Caemmerer, S. Biochemical Models of Leaf Photosynthesis; CSIRO Publishing: Collingwood, ON, Canada, 2000. [Google Scholar]
  67. Pfündel, E.E.; Klughammer, C.; Meister, A.; Cerovic, Z.G. Deriving fluorometer-specific values of relative PSI fluorescence intensity from quenching of F0 fluorescence in leaves of Arabidopsis thaliana and Zea mays. Photosynth. Res. 2013, 114, 189–206. [Google Scholar] [CrossRef] [PubMed]
  68. Moya, I.; Camenen, L.; Evain, S.; Goulas, Y.; Cerovic, Z.G.; Latouche, G.; Flexas, J.; Ounis, A. A new instrument for passive remote sensing: 1. measurements of sunlight-induced chlorophyll fluorescence. Remote Sens. Environ. 2004, 91, 186–197. [Google Scholar] [CrossRef]
  69. Magney, T.S.; Frankenberg, C.; Fisher, J.B.; Sun, Y.; North, G.B.; Davis, T.S.; Kornfeld, A.; Siebke, K. Connecting active to passive fluorescence with photosynthesis: A method for evaluating remote sensing measurements of Chl fluorescence. New Phytol. 2017, 215, 1594–1608. [Google Scholar] [CrossRef]
  70. Pfündel, E.E. Simultaneously measuring pulse-amplitude-modulated (PAM) chlorophyll fluorescence of leaves at wavelengths shorter and longer than 700 nm. Photosynth. Res. 2021, 147, 345–358. [Google Scholar] [CrossRef] [PubMed]
  71. Liu, X.J.; Liu, L.Y.; Hu, J.C.; Guo, J.; Du, S.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]
  72. Long, S.P.; Postl, W.; Bolhar-Nordenkampf, H.R. Quantum yields for uptake of carbon dioxide in C3 vascular plants of contrasting habitats and taxonomic groupings. Planta 1993, 189, 226–234. [Google Scholar] [CrossRef]
  73. Yin, X.; Struik, P.C. C3 and C4 photosynthesis models: An overview from the perspective of crop modelling. NJAS-Wagening. J. Life Sci. 2009, 57, 27–38. [Google Scholar] [CrossRef]
  74. Wu, A.; Doherty, A.; Farquhar, G.D.; Hammer, G.L. Simulating daily field crop canopy photosynthesis: An integrated software package. Funct. Plant Biol. 2017, 45, 362–377. [Google Scholar] [CrossRef] [PubMed]
  75. Egea, G.; Verhoef, A.; Vidale, P.L. Towards an improved and more flexible representation of water stress in coupled photosynthesis-stomatal conductance models. Agric. For. Meteorol. 2011, 151, 1370–1384. [Google Scholar] [CrossRef]
  76. Katul, G.; Manzoni, S.; Palmroth, S.; Oren, R. A stomatal optimization theory to describe the effects of atmospheric CO2 on leaf photosynthesis and transpiration. Ann. Bot. 2010, 105, 431–442. [Google Scholar] [CrossRef]
  77. Liu, Y.L.; Parolari, A.J.; Kumar, M.; Huang, C.W.; Katul, G.G.; Porporato, A. Increasing atmospheric humidity and CO2 concentration alleviate forest mortality risk. Proc. Natl. Acad. Sci. USA 2017, 114, 9918–9923. [Google Scholar] [CrossRef] [PubMed]
  78. Collatz, G.J.; Ball, J.T.; Grivet, C.; Berry, J.A. Physiological and environmental regulation of stomatal conductance, photosynthesis and transpiration: A model that includes a laminar boundary layer. Agric. For. Meteorol. 1991, 54, 107–136. [Google Scholar] [CrossRef]
  79. Ju, W.M.; Chen, J.M.; Black, T.A.; Barr, A.G.; Liu, J.; Chen, B.Z. Modelling multi-year coupled carbon and water fluxes in a boreal aspen forest. Agric. For. Meteorol. 2006, 140, 136–151. [Google Scholar] [CrossRef]
  80. Wu, A.; Hammer, G.L.; Doherty, A.; von Caemmerer, S.; Farquhar, G.D. Quantifying impacts of enhancing photosynthesis on crop yield. Nat. Plants 2019, 5, 380–388. [Google Scholar] [CrossRef]
  81. Wang, Y.P.; Leuning, R. A two-leaf model for canopy conductance, photosynthesis and partitioning of available energy Ι: Model description and comparison with a multi-layered model. Agric. For. Meteorol. 1998, 91, 89–111. [Google Scholar] [CrossRef]
  82. Porporato, A.; Laio, F.; Ridolfi, L.; Rodriguez-Iturbe, I. Plants in water-controlled ecosystems: Active role in hydrologic processes and response to water stress: III. Vegetation water stress. Adv. Water Resour. 2001, 24, 725–744. [Google Scholar] [CrossRef]
Figure 1. Representative images of the water-stressed winter wheat and variation in soil water content (θSWC, %) for the non-water stress (NS, blue) and water stress (WS, red) treatments under 20 days of progressive drought stress. The numbers above the images represent days after withholding water. Datapoints and error bars represent the mean ± standard deviation (SD) of four replicates.
Figure 1. Representative images of the water-stressed winter wheat and variation in soil water content (θSWC, %) for the non-water stress (NS, blue) and water stress (WS, red) treatments under 20 days of progressive drought stress. The numbers above the images represent days after withholding water. Datapoints and error bars represent the mean ± standard deviation (SD) of four replicates.
Plants 12 03365 g001
Figure 2. The impact of progressive drought stress on the responses of net photosynthetic carbon assimilation (Anet, μmol m−2 s−1), stomatal conductance (GS, mol m−2 s−1), actual rate of electron transport (Ja_PAM, μmol m−2 s−1), and full-band chlorophyll fluorescence emission at the photosystem level (ChlFP_F, μmol m−2 s−1) to changing light intensity. The number in the upper right corner of each plot is the number of days of progressive drought stress. The soil water content (θSWC, %) is also indicated. The solid lines represent the mean, and the shaded areas are ±1 standard deviation, of four replicates. Anet and GS are measured by the gas-exchange system.
Figure 2. The impact of progressive drought stress on the responses of net photosynthetic carbon assimilation (Anet, μmol m−2 s−1), stomatal conductance (GS, mol m−2 s−1), actual rate of electron transport (Ja_PAM, μmol m−2 s−1), and full-band chlorophyll fluorescence emission at the photosystem level (ChlFP_F, μmol m−2 s−1) to changing light intensity. The number in the upper right corner of each plot is the number of days of progressive drought stress. The soil water content (θSWC, %) is also indicated. The solid lines represent the mean, and the shaded areas are ±1 standard deviation, of four replicates. Anet and GS are measured by the gas-exchange system.
Plants 12 03365 g002
Figure 3. Relationships of net photosynthetic carbon assimilation (Anet, μmol m−2 s−1) with full-band chlorophyll fluorescence emission at the photosystem level (ChlFP_F, μmol m−2 s−1). Individual light–response curves are indicated by the black lines connecting measurements obtained at increasing light levels (0–2100 μmol m−2 s−1). Circles indicate no water stress (θSWC = 19.1%), diamonds represent moderate water stress (θSWC = 12.0%), and squares indicate plants under high water stress (θSWC = 9.7%). The number of light–response curves under high water stress was smaller than those under no water stress and moderate water stress due to human error, leading to unrealistic values in the measurements. Anet is obtained from the gas-exchange system.
Figure 3. Relationships of net photosynthetic carbon assimilation (Anet, μmol m−2 s−1) with full-band chlorophyll fluorescence emission at the photosystem level (ChlFP_F, μmol m−2 s−1). Individual light–response curves are indicated by the black lines connecting measurements obtained at increasing light levels (0–2100 μmol m−2 s−1). Circles indicate no water stress (θSWC = 19.1%), diamonds represent moderate water stress (θSWC = 12.0%), and squares indicate plants under high water stress (θSWC = 9.7%). The number of light–response curves under high water stress was smaller than those under no water stress and moderate water stress due to human error, leading to unrealistic values in the measurements. Anet is obtained from the gas-exchange system.
Plants 12 03365 g003
Figure 4. Light–response curves of the quantum yields (Ф) for the four different pathways during progressive drought stress: fluorescence (ФF, red), photosynthesis (ФP, green), regulated heat dissipation (ФN, yellow), and basal heat dissipation (ФD, blue). The number in the upper right corner of each plot indicates the day of progressive drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is also given. The inset panel illustrates the pattern of ФF under changing light intensity. The quantum yields (Ф) for the four different pathways are obtained from PAM measurements.
Figure 4. Light–response curves of the quantum yields (Ф) for the four different pathways during progressive drought stress: fluorescence (ФF, red), photosynthesis (ФP, green), regulated heat dissipation (ФN, yellow), and basal heat dissipation (ФD, blue). The number in the upper right corner of each plot indicates the day of progressive drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is also given. The inset panel illustrates the pattern of ФF under changing light intensity. The quantum yields (Ф) for the four different pathways are obtained from PAM measurements.
Plants 12 03365 g004
Figure 5. Relationships between the quantum yields of fluorescence (ФF) and photochemical quenching in PSII (red circle) during the progressive onset of drought stress. Polynomial models were used to fit all relationships (black solid lines). Breakpoints (black triangles) are the value of ФP at which the slope of the ФPF relationship changes sign.
Figure 5. Relationships between the quantum yields of fluorescence (ФF) and photochemical quenching in PSII (red circle) during the progressive onset of drought stress. Polynomial models were used to fit all relationships (black solid lines). Breakpoints (black triangles) are the value of ФP at which the slope of the ФPF relationship changes sign.
Plants 12 03365 g005
Figure 6. (a) Comparisons of measured versus simulated Vcmax (μmol m−2 s−1, Vcmax meas vs. Vcmax model) between 1 and 12 days after imposing drought stress; (b) comparisons of measured versus simulated Jmax (μmol m−2 s−1, Jmax meas vs. Jmax model) between 1 and 12 days after imposing drought stress. Datapoints and error bars represent the mean and standard deviation of four replicates.
Figure 6. (a) Comparisons of measured versus simulated Vcmax (μmol m−2 s−1, Vcmax meas vs. Vcmax model) between 1 and 12 days after imposing drought stress; (b) comparisons of measured versus simulated Jmax (μmol m−2 s−1, Jmax meas vs. Jmax model) between 1 and 12 days after imposing drought stress. Datapoints and error bars represent the mean and standard deviation of four replicates.
Plants 12 03365 g006
Figure 7. The measured versus simulated quantum yield of photochemical quenching in PSII (ФP) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is indicated. The number in the upper right corner of each plot indicates the day of progressive drought stress.
Figure 7. The measured versus simulated quantum yield of photochemical quenching in PSII (ФP) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is indicated. The number in the upper right corner of each plot indicates the day of progressive drought stress.
Plants 12 03365 g007
Figure 8. Measured versus simulated non-photochemical quenching (NPQ) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is shown. The day of progressive drought stress is indicated by the number in the upper right corner of each plot.
Figure 8. Measured versus simulated non-photochemical quenching (NPQ) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is shown. The day of progressive drought stress is indicated by the number in the upper right corner of each plot.
Plants 12 03365 g008
Figure 9. Measured versus simulated net photosynthetic carbon assimilation (Anet, μmol m−2 s−1) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is indicated. The number in the upper right corner of each plot indicates the day of progressive drought stress.
Figure 9. Measured versus simulated net photosynthetic carbon assimilation (Anet, μmol m−2 s−1) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is indicated. The number in the upper right corner of each plot indicates the day of progressive drought stress.
Plants 12 03365 g009
Figure 10. Measured versus simulated stomatal conductance (GS, mol m−2 s−1) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is shown. The day of progressive drought stress is indicated by the number in the upper right corner of each plot.
Figure 10. Measured versus simulated stomatal conductance (GS, mol m−2 s−1) during light–response curves between 1 and 12 days after imposing drought stress. The soil water content (θSWC, %) of progressive drought stress treatment is shown. The day of progressive drought stress is indicated by the number in the upper right corner of each plot.
Plants 12 03365 g010
Figure 11. Schematic of the leaf-scale concurrent light–response curve measurement system. The following measurements were taken: (a) pulse amplitude modulation (PAM) fluorometer (the quantum yields of photochemical quenching in photosystem II (ФP) and photochemically active radiation (PAR) are shown). (b) Gas exchange (net CO2 assimilation rate in light–response curves). (c) Chlorophyll fluorescence flux density emitted from photosystem II (ChlFP_F, μmol m−2 s−1), retrieved from the full ChlF emission spectrum. The external light source was attenuated by a customized 625 nm short-pass filter (not shown). A practical photograph of the leaf-scale concurrent measurement system which includes an LI-6800 gas-exchange chamber, a Dual-PAM-100 fluorometer, two HR2000+ spectrometers and two QE Pro spectrometers.
Figure 11. Schematic of the leaf-scale concurrent light–response curve measurement system. The following measurements were taken: (a) pulse amplitude modulation (PAM) fluorometer (the quantum yields of photochemical quenching in photosystem II (ФP) and photochemically active radiation (PAR) are shown). (b) Gas exchange (net CO2 assimilation rate in light–response curves). (c) Chlorophyll fluorescence flux density emitted from photosystem II (ChlFP_F, μmol m−2 s−1), retrieved from the full ChlF emission spectrum. The external light source was attenuated by a customized 625 nm short-pass filter (not shown). A practical photograph of the leaf-scale concurrent measurement system which includes an LI-6800 gas-exchange chamber, a Dual-PAM-100 fluorometer, two HR2000+ spectrometers and two QE Pro spectrometers.
Plants 12 03365 g011
Figure 12. Flowchart for quantifying the net CO2 assimilation rate Anet (μmol m−2 s−1) from the passive ChlF spectrum. The definitions of the variables are listed in Appendix A.
Figure 12. Flowchart for quantifying the net CO2 assimilation rate Anet (μmol m−2 s−1) from the passive ChlF spectrum. The definitions of the variables are listed in Appendix A.
Plants 12 03365 g012
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

Jia, Q.; Liu, Z.; Guo, C.; Wang, Y.; Yang, J.; Yu, Q.; Wang, J.; Zheng, F.; Lu, X. Relationship between Photosynthetic CO2 Assimilation and Chlorophyll Fluorescence for Winter Wheat under Water Stress. Plants 2023, 12, 3365. https://doi.org/10.3390/plants12193365

AMA Style

Jia Q, Liu Z, Guo C, Wang Y, Yang J, Yu Q, Wang J, Zheng F, Lu X. Relationship between Photosynthetic CO2 Assimilation and Chlorophyll Fluorescence for Winter Wheat under Water Stress. Plants. 2023; 12(19):3365. https://doi.org/10.3390/plants12193365

Chicago/Turabian Style

Jia, Qianlan, Zhunqiao Liu, Chenhui Guo, Yakai Wang, Jingjing Yang, Qiang Yu, Jing Wang, Fenli Zheng, and Xiaoliang Lu. 2023. "Relationship between Photosynthetic CO2 Assimilation and Chlorophyll Fluorescence for Winter Wheat under Water Stress" Plants 12, no. 19: 3365. https://doi.org/10.3390/plants12193365

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