Next Article in Journal
Analysis of the Food Loss and Waste Valorisation of Animal By-Products from the Retail Sector
Previous Article in Journal
Application of Machine Learning Models for Predictions on Cross-Border Merger and Acquisition Decisions with ESG Characteristics from an Ecosystem and Sustainable Development Perspective
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Light and Water Conditions Co-Regulated Stomata and Leaf Relative Uptake Rate (LRU) during Photosynthesis and COS Assimilation: A Meta-Analysis

1
Key Laboratory of Ecosystem Network Observation and Modeling, Institute of Geographic Sciences and Natural Resources Research, Chinese Academy of Sciences (CAS), Beijing 100101, China
2
School of Resources and Environment, University of Chinese Academy of Sciences (UCAS), Beijing 100049, China
3
School of Geography and Information Engineering, China University of Geosciences (Wuhan), Wuhan 430074, China
4
Hefei Institute of Physical Science, Chinese Academy of Sciences (CAS), Hefei 230031, China
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(5), 2840; https://doi.org/10.3390/su14052840
Submission received: 27 January 2022 / Revised: 21 February 2022 / Accepted: 22 February 2022 / Published: 28 February 2022

Abstract

:
As a trace gas involved in hydration during plant photosynthesis, carbonyl sulfide (COS) and its leaf relative uptake rate (LRU) is used to reduce the uncertainties in simulations of gross primary productivity (GPP). In this study, 101 independent observations were collected from 22 studies. We extracted the LRU, stomatal conductance (gs), canopy COS and carbon dioxide (CO2) fluxes, and relevant environmental conditions (i.e., light, temperature, and humidity), as well as the atmospheric COS and CO2 concentrations ( C a , COS and C a , CO 2 ). Although no evidence was found showing that gs regulates LRU, they responded in opposite ways to diurnal variations of environmental conditions in both mixed forests (LRU: Hedges’d = −0.901, LnRR = −0.189; gs: Hedges’d = 0.785, LnRR = 0.739) and croplands dominated by C3 plants (Hedges’d = −0.491, LnRR = −0.371; gs: Hedges’d = 1.066, LnRR = 0.322). In this process, the stomata play an important role in COS assimilation (R2 = 0.340, p = 0.020) and further influence the interrelationship of COS and CO2 fluxes (R2 = 0.650, p = 0.000). Slight increases in light intensity (R2 = 1, p = 0.002) and atmospheric drought (R2 = 0.885, p = 0.005) also decreased the LRU. The LRU saturation points of Ca,COS and Ca,CO2 were observed when ΔCa,COS ≈ 13 ppt (R2 = 0.580, p = 0.050) or ΔCa,CO2 ≈ −18 ppm (R2 = 0.970, p = 0.003). This study concluded that during plant photosynthesis and COS assimilation, light and water conditions co-regulated the stomata and LRU.

1. Introduction

1.1. Global Change and the Terrestrial Carbon Fixation

Global industrialization has caused the enrichment of atmospheric greenhouse gases (GHGs) globally [1]. In the atmosphere, GHGs have changed the process of radiation transmission [2]; further, more radiation is being trapped in the atmosphere [3,4]. The long-lived GHGs have dominated 47% of global warming, in particular, CO2 takes up 80% of this increase alone [5]. In this context, terrestrial and oceanic carbon sinks have been found to increase continuously over the last 60 years [6,7,8,9]. The terrestrial ecosystem responds to climate variability and climate change [10,11,12]. Global warming and land-cover changes have jointly accelerated carbon turnover [3,4]. The terrestrial carbon residence time decreased by one year in vegetation and by nine years in soil from the 1980s to the 2000s [13]. As the primary driver of terrestrial carbon accumulation or sequestration [14,15], GPP is not only the total carbon amount taken up by plants within photosynthesis [16] but also the largest CO2 flux in global carbon cycling [17]. Variations in temperature and precipitation, as well as extreme events, have affected GPP and the ecosystem respiration (ER) [18,19,20]. Moreover, in the terrestrial carbon–nitrogen cycle, such as CO2 fertilization [21], nitrogen deposition and other processes [22], a coupling interaction occurs with the systematic change of terrestrial carbon fixation [23]. Then, thickened tropospheric CO2 subtly influences GPP [24,25]. In network observations and ecological modeling studies, to represent the carbon that is actually fixed by the undisturbed biosphere through plant photosynthesis, the net ecosystem productivity (NEP) is a central concept in terrestrial carbon cycling [26]. NEP was originally defined as the imbalance between GPP and the ER [27]. Observations from networks of eddy covariance (EC) and footprint analysis provide rapid and accurate estimations of the exchange of mass and energy [28]. Furthermore, fluxes of CO2 and water (H2O) are indicators of ecosystem carbon and hydrological balances [29]. The EC micrometeorological technique, as well as the ecology-based biometric methods, are used to efficiently monitor the net ecosystem exchange (NEE) [30]. NEE is the net atmosphere–biosphere exchange of CO2 [31]. In the detailed mechanism, NEE is the sum of the ecosystem CO2 fluxes that do not involve the organic carbon either in its sinks or sources, as well as NEP [27]. Normally, NEE approximately equals to the negative NEP [26].

1.2. The Current State of Investigation on Photosynthesis and COS Assimilation

As a trace gas [32], the largest source of the COS atmospheric budget is the oceans of the Southern Hemisphere and the low latitude regions [33,34], as well as anoxic soils [35,36,37,38,39,40], i.e., directly from sulfur-containing organic matter or indirectly transferred from dimethyl sulfide (DMS) and carbon disulfide (CS2) [41,42,43,44]. Meanwhile, anthropogenic activities are the largest influencing factor of regional COS emission [45,46,47,48,49]. Among the whereabouts of atmospheric COS, vegetation photosynthesis plays a major role in tropospheric COS consumption [50]. From the process point of view, when catalyzed by the enzyme carbonic anhydrase (CA), COS can be irreversibly fixed in plants and soil ( COS + H 2 O HCOOS + H + ) [37,51,52,53,54,55,56]. In terms of the correlation between the photosynthetic hydrolysis rates of COS and CO2 [37,51,52,53,54,55,56] or the temporal dynamics of the proportional volume mixing ratio [32,41,42,57], the canopy assimilation rate of COS is correlated with that of CO2. Furthermore, unlike CO2, COS cannot be emitted through respiration. The COS flux is employed as an indicator of GPP in ecosystems [58,59,60]. Stomatal conductance (gs) can also be estimated by extrapolating the stomatal conductance of H2O (gs,H2O) from the stomatal conductance of COS (gs,COS) [61,62,63]. Based on laser spectroscopy technology [64], high-frequency observations of trace gases make it possible to measure the chamber-based or the EC flux COS [65,66,67]. At the regional scale, when using the processes-based land surface models (LSMs), such as SiB 4 and ORCHIDEE [68,69,70], to estimate the spatiotemporal variations of GPP [68,71,72,73], transpiration and gs [54,74,75], researchers take the atmospheric COS mixing ratio data of observation networks [58] and quantitative remote sensing as the input variables [76,77,78,79,80]. Global studies take the carbon assimilation systems, such as Carbon Tracker-Langrange, to recognize the sources and sinks of COS [71,81]. Recently, scientists are also making efforts to identify the missing source and sinks of global atmospheric COS [60,82,83,84]. It is urgent to precisely formulate the regulation mechanism of canopy COS assimilation, as well as the linkage between canopy COS flux (FC,COS) and GPP. The following field observations have studied the relationship between environmental factors and COS assimilation. Site observation of bryophytes found that the biological fixation of COS is mostly regulated by meteorological factors [85]. In plant leaf photosynthesis, during hydrolysis, the catalysis of COS by the CA enzyme is light-independent [37,41] and more efficient than that of CO2 [86]. During the light response of photosynthesis, CO2 is preferentially catalyzed by RuBisCO over COS by a factor of 110 [51,87]. Moreover, the catalytic efficiency of the CA enzyme, which is highly abundant, exceeds that of RuBisCO overall [37,86,88,89]. At the leaf scale, stomata are not only the gateway for photosynthetic CO2 assimilation but also the vapor flux in transpiration. At the same time, COS is assimilated through the diffusive pathway (e.g., the leaf boundary layer, stomata and mesophyll). From the process perspective, environmental conditions (e.g., light, temperature, leaf moisture conditions and the CO2 concentration in the leaf boundary) have an important role in the conductance of both the stomata and mesophyll [90]. In response to leaf drought, to protect the mesophyll cells from excessive water loss, abscisic acid (ABA) is transferred to the stomata by the catheric sap flow [91], which stimulates the striction of guard cells and the stomata close. The gradient between the atmospheric CO2 concentration (Ca,CO2) and the intercellular CO2 concentration (Ci,CO2) is mainly regulated by the stomatal guard cells. Specifically, the mesophyll CO2 concentration is controlled by the diffusive pathway. Then, the above processes regulate CO2, COS and H2O fluxes [92,93]. There are differences between the plant morphology (herbs, trees), tree foliage types (conifer, deciduous and evergreen) and the photosynthesis pathways (C3 and C4 plants). It has been estimated that 85% of photosynthetic CO2 uptake is regulated by gs and mesophyll conductance (gm) [94]. Stomata play an important role in linking carbon and water processes [95]. On the other hand, because COS is a trace gas in the troposphere and the catalytic efficiency of CA enzyme is high, the intercellular COS concentration (Ci,COS) is kept at a very low level [41,51,96]. As they share the same assimilation pathway, COS and CO2 are also maximally resisted by the stomata and the mesophyll [37,73,97]. Furthermore, based on parallel observations of the atmospheric concentration of COS (Ca,COS) and Ca,CO2 [56,98], recent studies found that COS can be used to obtain gs across different biomes and scales [60,73,99]. From the perspective of both diel and seasonal variation, COS can not only be slightly continuously taken up by the partially closing stomata at night [65], but when some branches wither, COS can be released from these dead or drought-damaged components [65,100,101]. In the process of plant photosynthesis, by eliminating the interference of non-foliage components [90], the conductance of the leaf boundary (gb) is caused by adhesion and COS. The COS assimilation follows F c , COS = ( 1 / g b + 1 / g s + 1 / g m ) 1 · ( C a , COS C i , COS ) [51,56,60,73,97]. The pathway conductance controls the COS assimilation rate at the leaf scale [60]. Meanwhile, the COS flux is found to be regulated by radiation and the vapor pressure deficit (VPD) [102,103,104,105]. In the diffusional pathway, environmental conditions, such as radiation, the leaf VPD, relative humidity (rH), nitrogen deposition, salinity and aerosol concentrations, also affect stomata conductance [106,107,108]. In the process of COS, CO2 and H2O exchange, the molecule of COS is larger than CO2 and the adhesional coefficient of COS is different from CO2 [51,56]. Moreover, stomata strongly respond to COS, which is mediated by H2S [51]. Further, in hydrolysis, the mesophyll COS is catalyzed by the CA enzyme [37,41], while in photoreaction, CO2 is catalyzed by RuBisCO [51,87]. This causes a different diffusional coefficient between COS and CO2 in each pathway. For other processes, because COS cannot be released by plant respiration, the difference is also shown in the photosynthetic feedback pathway between COS and CO2 [109]. The assimilation processes of both COS and CO2 also vary between species. It is important to explore the differential regulation mechanism of environmental conditions in the whole diffusional pathway of COS and CO2 assimilation (Figure 1). In observational and modeling studies, water use efficiency (WUE), as the ratio of carbon assimilation (i.e., GPP) to water losses, has been employed to represent the coupling of the carbon and water cycles. Controlled by leaf stomata, WUE is broadly correlated with Ca,CO2 at the leaf boundary layer [110].

1.3. Background and Main Goals of Investigations

As an improvement of the traditional partition method of NEE [111], the leaf relative uptake rate ( LRU = F COS × C a , CO 2 GPP × C a , COS ) has been proposed for use in estimating GPP from the COS flux [51,56,59,96,112]. LRU has been estimated to be in the range of 1.5~4.0 for many species [51,54,56,113]. On the other hand, the LRU is also an ecophysiological indicator used in estimating Ci,CO2/Ca,CO2 and gs [56,90]. COS assimilation of different plant components [51], litters and soil responses vary distinctly during at diel and seasonal scales, leading to a fluctuation in the LRU value [90]. Further, the differential behavior of COS and CO2 in physiological and meteorological processes causes small fluctuations in the plant assimilation ratio of the two, which further affects the LRU value [51,101,114,115]. Generally, LRU is controlled by gs, as well as concentrations of either COS or CO2 in the leaf boundary layer [54]. The environmental conditions, atmosphere and vegetation at certain stations may influence the LRU [51,65,88,116]. In the species-dependent LRU environment response process [65,101,113,114], light intensity, temperature and water limitation [115], as well as atmospheric conditions, such as the frictional wind speed, Ca,COS and Ca,CO2, are influential [51,68]. In this study, we hypothesized that regarding diurnal, daily and seasonal variation, the above environmental conditions (i.e., light, temperature, water conditions, Ca,COS and Ca,CO2) co-regulate gs, photosynthesis and COS assimilation, and this species-dependent relationship will affect the LRU (Figure 1). By providing perspectives on the development of COS-constrained GPP simulations and the key physiological processes within the LRU regulation mechanism, this study aimed to determine the relationship between environmental conditions and LRU variations, to describe how stomata regulate the relationship between FC,COS and canopy CO2 flux (FC,CO2) and to constrain the accuracy of LRU estimates.
Figure 1. Schematic diagram of COS and CO2 assimilation during leaf photosynthesis. The diffusion processes of COS and CO2 are generalized from Section 1.2. The current state of investigation on photosynthesis and COS assimilation by us. Blue font characters are the key environmental conditions we filtered from the literature. In the corresponding physiological processes, they act as regulators. The source literature is as follows: light intensity [37,41,51,86,87,90], rH [61,62,63,85,90,91,92,93,94,106,107,108,115], VPD [102,103,104,105] and Ca,COS and Ca,CO2 [37,41,50,51,52,53,54,55,56,60,73,87,97,98,109].
Figure 1. Schematic diagram of COS and CO2 assimilation during leaf photosynthesis. The diffusion processes of COS and CO2 are generalized from Section 1.2. The current state of investigation on photosynthesis and COS assimilation by us. Blue font characters are the key environmental conditions we filtered from the literature. In the corresponding physiological processes, they act as regulators. The source literature is as follows: light intensity [37,41,51,86,87,90], rH [61,62,63,85,90,91,92,93,94,106,107,108,115], VPD [102,103,104,105] and Ca,COS and Ca,CO2 [37,41,50,51,52,53,54,55,56,60,73,87,97,98,109].
Sustainability 14 02840 g001

2. Materials and Methods

2.1. Data Preparation

We used F C , COS , F C , CO 2 and the relevant environmental conditions to explain the mechanism of COS assimilation during photosynthesis and the LRU regulation mechanism based on stomata. During the data preparation, we collected studies published during 1985~2020 from Google Scholar and Web of Science, which simultaneously observed gs, FC,COS, FC,CO2, environmental conditions and the simulated LRU [14,51,65,90,96,101,113,114,115,117,118,119,120,121,122,123,124,125,126,127,128,129]. According to the PRISMA statement of systematic reviews and meta-analyses, when retrieving these studies, we used measurement, photosynthesis and gs as the keywords, and then selected 228 studies. Then, we used measurement, photosynthesis and COS as the keywords, and then selected 58 studies. In the intersection of the two folders there were 25 studies that simultaneously observed photosynthesis, COS and gs. Then, to constrain the uncertainty caused by the temporal variation of observations from instruments and the parameters, we paid no consideration to the belowground parts [90] and just chose measurements of leaves. Ultimately, 22 datasets from 15 sites were chosen for this meta-analysis. During measurements of the leaf gas exchange and g s , these studies also measured relevant environmental conditions. We collected four categories of environmental conditions, and light conditions were represented by light intensity (μmol m−2 s−1) or photosynthetic active radiation (PAR, μmol m−2 s−1), temperature conditions were represented by the canopy or soil temperature (°C) and water conditions were represented by rH (%) and VPD (Pa). Meanwhile, the atmospheric COS and CO2 concentrations were represented by Ca,COS (ppt) and Ca,CO2 (ppm).
Raw data were acquired from the tables, figures and appendixes of the publications using Engauge Digitizer (Free Software Foundation, Inc., Boston, MA, USA). Relationships between the environmental conditions and dependent variables (i.e., LRU, gs, FC,COS, and FC,CO2) identified in the studies were also collected for verification of the results. We filtered the search results to find studies performed over the plant photosynthetic part of terrestrial ecosystems and classified the research in previous studies on the basis of their study sites. Then, these studies were categorized into subsets of biomes on the basis of their study sites. These biomes included needle-leaf forest (NLF), broadleaf forest (BLF), mixed forest (MXF), grassland (GRA), marsh (MSH), cropland dominated by C3 plants (CPLC3) and cropland dominated by C4 plants (CPLC4) (Figure 2).
In each biome group, the studies were separated according to their measurement frequencies. The data from studies that included data from diurnal variations were classified into groups as noon vs. morning and evening (N vs. ME), day vs. night (D vs. N) and forenoon and afternoon vs. dawn and dusk (FA vs. DD) data; FA vs. DD was included to avoid the influence of inverted light-impacted stomatal closures at midday [130]. Data from periods longer a day, such as daily and seasonal records, were employed only for testing. The summer solstice day and the autumnal equinox day (normally on 20 June and 23 September, DOY 171 and 266) were used as the beginning and end of summer, respectively. For comparison, PAR = 600 μmol m−2 s−1 was the criterion of sunny and cloudy conditions when separating the records of the control experiments (Table 1). Notably, different biomes in the same study were classified as separate groups, and data from the same study site recorded at different temporal scales were separated into independent groups. A total of 101 separate studies were used to establish the dataset. The mean value of each variable and its sampling size and standard deviation were also collected. The sample size of each group and each indicator, as well as the characteristics of the variations in environmental conditions, are listed in Table S1.

2.2. Meta-Analysis

For different dependent variables, the changes in environmental conditions were quantified between the control and the treatment. To estimate the impact of environmental condition changes between the controls and treatments, statistical parameters (e.g., the count ( N ¯ ), average value ( X ¯ ) and standard deviation ( S D ¯ )) for both the controls and treatments were used to perform meta-analysis for each group with METAWIN 2.1 [131]. Then, the effect size (d) was calculated as follows [73]:
d = N ¯ E N ¯ c s J
s = ( n E 1 ) ( s E 2 ) + ( n c 1 ) ( s C 2 ) n E + n c 2
where the subscripts E and C denote experimental and control groups, respectively; N ¯ i is the mean; ni is the sample size; and s i 2 is the variance for the experimental (i = E) or control (i = C) groups; J is used to correct for bias due to small sample size [73].
A response ratio (RR) was also calculated as follows [132]:
RR = ln ( X t ¯ X c ¯ ) = ln ( X t ¯ ) ln ( X c ¯ )
where X t ¯ and X c ¯ are the means of a variable in the treatment (Tables S2 and S3) group and the control group, respectively [132]. Its variance ( v ) was estimated as follows:
v = S c 2 n c X c 2 ¯ + S t 2 n t X t 2 ¯
where n c and n t are the sample sizes in the control and treatment groups, respectively; Sc and St are the standard deviations in the control and treatment groups, respectively.
The impact of the change in environmental conditions on Hedges’ d or ln ( RR ) was considered significant if the 95% confidence interval (CI) did not include zero. The total heterogeneity ( Q total ) of each group was classified into within-group heterogeneity ( Q within ) and between-group heterogeneity ( Q between ). If the probability value Q between 0.05 , the difference between the groups or biomes was considered significant. In detail, Q b , Hedges d is the Q between value of Hedges d , and when the groups are fixed to biomes, Q b , Hedges d , biomes was used. The response function that fitted the data from the most studies was chosen to represent the dependent variables’ responses to the changing environmental conditions. The mean effect size, 95% CI and sample number were calculated for each biome and group with MATLAB R2016b (The MathWorks, Inc., Natick, MA, USA).

2.3. Statistical Analysis

To investigate the changes in environmental conditions regulation of stomata and the responses of FC,COS and FC,CO2, the relationships between FC,COS and variation in Ca,COS or Ca,CO2 and the relationships between FC,CO2 and variation in Ca,COS or Ca,CO2 were compared. These comparisons further clarified whether the LRU would be adjusted along with the variation in gs. First, we chose groups set by diurnal variations and compared the linear, quadratic or cubic regression between the Hedges’ d of each dependent variable and the change in each environmental condition, as well as its reciprocal transformation. Significant intercorrelations were obtained to explain the impact of the change in environmental conditions on each dependent variable. Second, we performed the aforementioned regression between the Hedges’ d value of the LRU, gS, FC,COS and FC,CO2 in pairs to clarify the roles of FC,COS and FC,CO2 regarding how changes in environmental conditions regulate gS and the synchronous variation in the LRU; these results corroborated the correlation between FC,COS and FC,CO2, as well as how the LRU responded to the stomata. Scatter plots with regression lines were created with SigmaPlot 12.5 (Systat Software, Inc., San Jose, CA, USA).

3. Results

3.1. Changes in Environmental Conditions

Among the diurnal variation groups, in each LRU, gs, FC,COS and FC,CO2 group, the light intensity was increased in 100%, 73%, 89% and 90% of the studies, respectively, in each group (Figure 3a). Moreover, 89%, 63%, 71% and 69% of the studies included a warming process, respectively (Figure 3b). Although characterized by VPD, the water conditions at the canopy boundary were increased in 100% of the overall studies, and the rH values showed that the atmospheric moisture was greatly diminished in 75%, 50%, 71% and 60% of the studies, respectively (Figure 3c,d). For Ca,COS and Ca,CO2, 90%, 50%, 74% and 71% of the studies showed enhanced Ca,COS, respectively (Figure 3e). In contrast, 86%, 50%, 36% and 42% of the studies showed diminished Ca,CO2, respectively (Figure 3f).

3.2. Changes in the Leaf LRU, gs, FC,COS and FC,CO2

The meta-analyses showed that there were significant differences in the LRU between the groups categorized by time scale, which reflected the tendency of environmental conditions ( Q b , Hedges d   <   0.01 ). Furthermore, no significant discrepancy was observed between the groups at the same time scales ( Q b , Hedges d = 0.56 ) (Figure 4a). Under the aforementioned environmental condition changes, the LRU decreased in the “Total”, “Control” and the N vs. ME groups among the groups categorized by time scale, which was significantly stronger in the NLF, MXF and CPLC3 biomes. Solely on the aspect of light intensity, the decrease of LRU was noticed in the lower light “control” group. Moreover, no significantly different LRU responses among biomes were obtained ( Q b , Hedges d , biomes = 0.12) in response to environmental condition changes. A change in the LRU was associated with enhanced gs due to stomatal opening in all the groups, except the daily time scale group. Although no significant differences were found between biomes ( Q b , Hedges d , biomes = 0.73 ), the gs in MXF, CPLC3 and CPLC4 showed the strongest response to environmental condition changes (Figure 4b). For the photosynthetic gas fluxes, no significantly different FC,COS responses to environmental condition changes were found between the groups categorized by time scale ( Q b , Hedges d = 0.44 ). After experiencing the strongest variation of VPD, Ca,COS and Ca,CO2 (Figure 3d–f), FC,COS was diminished in most samples of N vs. ME, FA vs. DD, D vs. N and “Total” groups; meanwhile, only in N vs. ME was this decrease significant. The diminished FC,COS was also found in the NLF, BLF, GRA and CPLC3 biomes. In contrast, in the biomes of MXF, MSH and CPLC4, FC,COS was found to have increased. At the same time, significantly different FC,COS responses were found between biomes ( Q b , Hedges d , biomes = 0.00 ) (Figure 4c). Moreover, no significant differences in the FC,CO2 response were found between the different temporal groups ( Q b , Hedges d = 0.74 ). FC,CO2 decreased significantly in all the N vs. ME, FA vs. DD, D vs. N and “Total” groups, which reflected FC,CO2 diurnal variations. Significantly different FC,CO2 responses were found between biomes ( Q b , Hedges d , biomes = 0.00 ). In NLF, BLF, GRA and CPLC3, the environmental conditions had an evident negative impact on FC,CO2. However, FC,CO2 increased in MXF and MSH under the same environmental conditions variation (Figure 4d).

3.3. The Dependent Variables as Regulated by Environmental Conditions

On the groups categorized by diurnal cycling (N vs. ME, FA vs. DD, D vs. N), the dependent variables (such as LRU, gs, FC,COS and FC,CO2) were regulated by the environmental conditions (light intensity, temperature, rH, VPD, Ca,COS and Ca,CO2). Of the environmental conditions, the LRU was significantly influenced by variations in light intensity (R2 = 1, p = 0.002), VPD (R2 = 0.885, p = 0.005), Ca,COS (R2 = 0.58, p = 0.05) and Ca,CO2 (R2 = 0.97, p = 0.003) (Figure 5a–f). There was an exponential relation between light intensity and the LRU (Figure 5a). On the other hand, increased VPD reduced the LRU (Figure 5d). As Ca,COS increased, the LRU increased until a certain turning point of Ca,COS and then decreased. The regression suggested the LRU saturation point could be ΔCa,COS ≈ 13 ppt (Figure 5e). In contrast, between LRU and ΔCa,CO2, the LRU saturation point could also be ΔCa,CO2 ≈ −18 ppm (Figure 5f).
As shown by the linear regression of the change in environmental conditions to gs, the stomata were opened significantly only by light intensity (R2 = 0.58, p = 0.01) (Figure 6a).
The fitted FC,COS increased significantly with a decrease in light intensity and was significantly decreased with a slight increase in light intensity (R2 = 0.23, p = 0.01) (Figure 7a,b).
In addition, FC,CO2 decreased linearly with increasing temperature (R2 = 0.80, p = 0.00) (Figure 8b). In terms of moisture conditions, FC,CO2 increased linearly with increasing rH (R2 = 0.83, p = 0.00) (Figure 8c). However, at the canopy boundary layer, the change in VPD did not show any significant correlation with either flux (Figure 7d and Figure 8d). From the perspective of gas exchange, FC,COS was not significantly limited by Ca,COS (Figure 7e). Nevertheless, there was a compensation point for Ca,CO2 assimilation (ΔCa,CO2 ≈ 10 to 40 μmol m−2 s−1), under which, FC,COS was the weakest (R2 = 0.30, p = 0.04) (Figure 7f). FC,CO2 decreased linearly with increased Ca,COS (R2 = 0.19, p = 0.04) but increased linearly with increased Ca,CO2 (R2 = 0.19, p = 0.04) (Figure 8e,f).

3.4. The Role of FC,COS and FC,CO2 in the Mechanism of LRU and gs Regulation

In the diurnal cycling groups, there was no significant correlation between either the Hedges d of FC,COS or the Hedges d of FC,CO2 with the Hedges d of the LRU (Figure 9a,b).
However, FC,COS was significantly correlated with gs (R2 = 0.34, p = 0.02) (Figure 10a). The Hedges d of gs exhibited a compensation point at a certain level of the Hedges d of FC,COS (Figure 10a).
The linear correlation between the Hedges d of FC,COS and the Hedges d of FC,CO2 was determined (R2 = 0.65, p = 0.00) (Figure 11a). Ultimately, when doing correlation analysis, there were not enough samples to separate the control group or the groups of CPLC3, CPLC4 and MXF from the total. Surprisingly, there was no significant correlation between the Hedges d of gs and the Hedges d of the LRU (Figure 11b).

4. Discussion

4.1. Co-Response of LRU and Stomata to the Variations in Environmental Conditions

The changes in environmental conditions under diurnal cycling could be characterized as increases in light and temperature. Drought was characterized by a decrease in rH or an increase in VPD. For most studies investigating either LRU or FC,COS, Ca,COS and Ca,CO2 increased, while for most studies that investigated FC,CO2, Ca,CO2 decreased.
The gs models (i.e., the Ball–Berry model) already simplified the regulating effects of environmental conditions on gs at the leaf scale, and the “big leaf” model cannot be directly applied at the canopy scale [133]. However, this study indicated that under increasing light, the LRU response may be the opposite of the increase in gs, even though there is no evidence to illustrate the correlation between gs and the LRU. Under the change in environmental conditions, significantly opposite responses of the LRU and gs were found, mostly in the MXF and CPLC3 biomes; the significance was confirmed by comparison with the control groups. Only in the MXF, CPLC3 and CPLC4 biomes was gs increased. Moreover, an increase in gs with a decrease in the LRU was found in the N vs. ME group and the controls. Due to the lack of data, the LRU regulation mechanism in the GRA and CPLC4 biomes was unclear. Due to the lack of separation between evergreen and deciduous BLF and the strong influence of moist-oxic soil [43], the correlations between the LRU and gs in the NLF, BLF and MSH biomes were nonsignificant.
From the response of LRU to the change in environmental conditions, the decrease in the LRU with the increase in light and VPD found in this study was consistent with the findings of [115]. During an increase in light intensity, the decrease rate of the LRU slowed, as the LRU became saturated when exposed to strong light [115]. Moreover, the saturation point was greater than 380 μmol m−2 s−1. At the canopy scale, the main factors that are synergetic with gs include the diffuse light fraction and VPD [73,121,134], which showed a trade-off relationship with the LRU. LAI influences gs at the canopy scale as well [73,121,134]. Stomata are typically open under elevated Ca,COS conditions [51,124], and in this study, the LRU reached peak values when ΔCa,COS ≈ 13 ppt, as well as when ΔCa,CO2 ≈ −18 ppm.

4.2. Coresponse of Canopy COS and CO2 Fluxes to Variations in Environmental Conditions

At timescales under diurnal cycling, FC,COS and FC,CO2 both decreased in FA vs. DD and D vs. N under the changes in environmental conditions. However, FC,COS increased in the MXF, CPLC3, and CPLC4 biomes, which was also found in most studies of NLF. In addition, in the N vs. ME group, FC,CO2 in MXF and MSH increased in the FA vs. DD and D vs. N groups, and FC,CO2 decreased in NLF, BLF, GRA, and CPLC3.
In this study, when the light intensity was reduced by 300~700 μmol m−2 s−1, FC,COS was composed of continuing assimilation by the leaves. For instance, stomata open in the dark and cause COS assimilation [135], which makes up 17% of daily FC,COS [118]. Under a minor increase in light intensity, FC,COS significantly decreased. However, FC,CO2 increased under cooler conditions. In croplands and mid-latitude forests, FC,COS also decreased under water shortage [65,101,136]. Water shortage, an increase in temperature, an increase in Ca,COS and a decrease in Ca,CO2 were all main factors resulting in a decrease in FC,CO2, as the tissue water content is important in canopy CO2 assimilation [137]. However, an slightly increase of Ca,CO2 value also indicated the lowest FC,COS, as there was a hysteresis between FC,COS and FC,CO2 fluctuations within diurnal cycles [96]. Simultaneous increases in FC,COS and FC,CO2 were found only in MXF. In contrast, in NLF and CPLC3, only FC,CO2 decreased. In the nonstomatal structures, the COS assimilation rate increased with water shortage [138,139,140]. For example, in some bryophytes, mosses, liverworts and lichens, as the hydrolysis process is catalyzed by the light-independent CA enzyme [37] and is affected by reverse hydrolysis, the net absorption rate of COS is independent of light conditions [85]. Bidirectional hydrolysis has also been observed in some fungi and bacteria [141,142,143]. This retrohydrolysis does not cause COS emissions and influence the FC,COS of the higher plants overall [51,56,144]. For higher plants, although there is a reverse hydrolysis process in the stomata opening process, resistance and osmotic pressure stopped COS from being emitted; the light-independent CA enzyme is still the key to the absorption rate of COS during photosynthesis [145,146]. Enhanced frictional wind at sunrise caused COS emission from the canopy gap; furthermore, soil microorganism breakdown boosted by light seems to be the reason for the significantly increased FC,COS. Recent studies also discussed whether FC,COS would be affected by the diversity of the enzyme CA between different species [147,148]. When the COS concentration in the subcavities of the mesophyll (Cm,COS) is close to zero, FC,COS is strongly affected by Cm,COS [144,149]; however, no correlation was found between the changes in Ca,COS and FC,COS.

4.3. The LRU Variation Mediated by the Canopy COS and CO2 Fluxes under the Regulation of Stomata

Most of the gas exchange in leaflets is controlled by the stomata [73,150]. In this study, at the canopy scale, no evidence was found regarding the correlation between the LRU and FC,COS or FC,CO2. In the diurnal cycling groups, increased gs and decreased FC,COS were found to occur concurrently, except in the case of N vs. ME. This difference in N vs. ME could have been due to the stomatal closure that occurs at midday [130]. However, in non-wetting vegetation, such as NLF, MXF, CPLC3 and CPLC4, moderate stomatal closure and increased enzyme CA activity caused an increase in the FC,COS of N vs. ME. The LRU and gs were found to respond in opposite ways to the changes in environmental conditions in MXF and CPLC3. Moreover, the LRU was found to differ between species. In the NLF and CPLC3 groups, the LRU decreased with increasing FC,COS and decreasing FC,CO2. In contrast, increases in both FC,COS and FC,CO2 were found in GRA in which the LRU was found to decrease.
In the process of changing environmental conditions, under a minor increase in light intensity, FC,COS decreased. Furthermore, synchronization was observed between the decrease in FC,CO2 caused by warming and the decrease in the LRU caused by an increase in light. In contrast, gs increased when the light intensity increased. Recent studies noted that pathway conductance and CA enzyme activity colimited FC,COS [73]. Furthermore, the response of stomata to temperature was not very sensitive [151], as the CO2 concentration in the subcavities of the mesophyll and the activity of enzyme CA colimited conductance throughout the whole pathway [138]. This study found a linear trade-off relationship only between FC,CO2 and temperature. However, there was no correlation between FC,CO2 and gs because GPP and ER were not partitioned from FC,CO2. Meanwhile, water shortage resulted in decreased FC,CO2 (consistent with the decreases in rH and VPD).
In the diurnal cycling groups, there was a synergetic correlation between FC,CO2 and the variation in the rH or Ca,CO2. However, FC,CO2 exhibited a trade-off with temperature or Ca,COS. Then, stomata closure was controlled by the Ca,COS decrease [51,124]. Moreover, enhanced friction wind caused COS emission from the canopy and induced a quadratic correlation between Ca,CO2 and FC,COS. Therefore, the variation in FC,COS was smaller than that in FC,CO2 [96]. However, no significant correlation was found between the LRU and either FC,COS or FC,CO2. Environmental conditions such as light intensity, VPD, Ca,COS and Ca,CO2 were the key factors influencing LRU variation. As evidenced by the results from NLF, MXF and CPLC3, Ca,COS and Ca,CO2 increased in the diurnal cycling groups in non-wetting biomes. During the diurnal variation, the canopy COS absorption flux mainly occurred in the daytime. First of all, a minor increase in light intensity enhanced gs, and enhanced friction wind caused COS emission from the canopy gap; furthermore, FC,COS decreased. Furthermore, as the intercomparison of global studies revealed, as Ca,CO2 increased by 10 to 40 μmol m−2 s−1, FC,COS decreased to its lowest and rebounded. The LRU was influenced by both Ca,COS and Ca,CO2. In contrast, as the stomata closed, the COS diffusion pathway was inhibited due to water shortage at midday; even if Ca,COS and Ca,CO2 increased, FC,COS was still decreased. Finally, FC,COS was the lowest when the stomatal openings were small. Notably, the increase and decrease in FC,COS could both be related to an increase in gs. The regulation mechanism is important in formulating the LRU time series and in upscaling the LRU into the ecosystem relative uptake (ERU) and the atmospheric relative uptake (ARU) [122]. However, the nonsignificant relationship between the Hedges’ d of gs and the Hedges’ d of the LRU limited our ability to illustrate how the variations in environmental conditions regulate stomata and fluxes and thus further influence the LRU.

4.4. Employing the LRU to Constrain the Uncertainties in COS-Based GPP Simulations

A potential improvement to the partitioning of NEE [152] is based on the function relating FC,COS, the GPP and the LRU [43,122]. The quantum cascade is an advanced technique employed in the continuous measurement of FC,COS; in this system, R d is much smaller than in other systems [54,56,153]. However, during the COS-based simulation of the GPP, COS emissions also occur in dehydrating leaves [85], which would lead to systematic noise. Moreover, the partial nocturnal closure of stomata, which is the reason for the depletion of the nighttime Ca,COS [65,73,101,113,114,129,154,155], was also observed in certain species [101,113]. Overall, at night, even though COS hydrational assimilation on the water surface and in oxic soils, as well as oxidative release in anoxic soils [141], participate in COS circulation, the terrestrial nighttime FC,COS is mostly influenced by the plant canopy [65,73,113,129,156]. It is also worth quantifying the influence of COS nighttime flux on the LRU at the diurnal scale. Recent studies have drawn attention to the fluctuation of FC,COS [73,114]. These studies noted that even though FC,COS was considered to be regulated by the activity of the enzyme CA and the LAI of the plant canopy, both diurnal and seasonal fluctuations in Ca,COS and Ca,CO2 were correlated with FC,COS [73,114]. Additionally, Ca,COS and Ca,CO2 peaked at night when FC,COS was low [157,158]. During the heavy turbulence at dawn, as COS and CO2 were replenished in the canopy and subcanopy airspace, the canopy storage effect also influenced the simulated FC,COS and FC,CO2 [73]. It is also worth exploring whether this process influences the LRU.
Further studies should be performed to identify the additional key factors that regulate LRU variation [159] and the influence of NEE partitioning. Layered observations led to a proposal to partition FC,COS into leaf assimilation and soil exchange components [42]. Moreover, soil COS consumption, which is linearly correlated with soil temperature and makes up 34~40% of the nighttime FC,COS in the boreal forest, reached its saturation point at 20% water content [118,156]. It is worth further exploring the considerable differences between the environmental conditions regulation mechanism for COS leaf assimilation and that for COS soil exchange.

5. Conclusions

During plant photosynthesis and COS assimilation, light and water conditions co-regulated the stomata and LRU. Overall, at diurnal scales, FC,COS was controlled by the stomata and was decreased by water shortage in non-wetting biomes; at the same time, the leaves were exposed to increased Ca,COS. However, under variations in the environmental conditions, a three-step response occurred in FC,COS. First, slightly increased light increased the gs, as well as the COS emission from the canopy gap; then, FC,COS was decreased. Second, when the stomata were opening, since the subcanopy emission of COS and the enzyme CA catalysis rate of COS exceed CO2, a Ca,CO2 increase of 10 to 40 μmol m−2 s−1 was a compensation point for COS assimilation, at which point, FC,COS was the lowest. Third, the stomata closed under both high light intensity and high temperatures, and FC,COS was decreased as well. Although this study provided no evidence of the correlation between the LRU and gs, they were shown to be regulated by the variations in environmental conditions in opposite ways, at least in MXF and CPLC3. Moreover, the LRU decreased under slightly increased light and atmospheric drought. The LRU saturation points of Ca,COS and Ca,CO2 were observed when ΔCa,COS ≈ 13 ppt or ΔCa,CO2 ≈ −18 ppm. In further studies, we will formulate the quantitative model of the LRU regulation mechanism based on the above-mentioned conclusion using gs, Ca,COS, Ca,CO2, light and water conditions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su14052840/s1, Table S1: The changes in environmental conditions as well as the variations of Hedges’d and Ln(RR) among different dependent variables (e.g., LRU, gs, FC,COS, FC,CO2) and different groups; Table S2: The studies collected in this meta-analysis; Table S3: Dependent variables (e.g., LRU, gs, FC,COS, FC,CO2) and related environmental conditions collected in each study; Section S1: Studies collected in this meta-analysis; Section S2: The abbreviations used in this study.

Author Contributions

Conceptualization, S.W., L.W. and J.C.; Data curation, P.W. and J.C.; Formal analysis, P.W.; Investigation, P.W., L.W., J.C., L.M. and X.W.; Methodology, P.W., L.M. and X.W.; Project administration, S.W. and L.W.; Resources, P.W. and K.Z.; Software, P.W., X.W. and Y.L.; Validation, B.C., L.M. and Y.L.; Visualization, P.W., Y.L. and K.Z.; Writing—original draft, P.W.; Writing—review and editing, S.W., B.C., M.A. and J.C. All authors have read and agreed to the published version of the manuscript.

Funding

This work was finally supported by the Scientific Instrument Developing Project of the Chinese Academy of Sciences, Grant No. YJKYYQ20190042, and the National Key Research and Development Program of China No. 2017YFC0503803.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data collected in this article is within the “Supplementary Materials” document.

Acknowledgments

We are grateful to the scientists who contributed to the global database used in this meta-analysis.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Change, C. Working Group I: The Scientific Basis. In Third Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University: New York, NY, USA, 2001. [Google Scholar]
  2. Hertzberg, M.; Siddons, A.; Schreuder, H. Role of greenhouse gases in climate change. Energy Environ. 2017, 28, 530–539. [Google Scholar] [CrossRef] [Green Version]
  3. Piao, S.; Wang, X.; Park, T.; Chen, C.; Lian, X.; He, Y.; Bjerke, J.; Chen, A.; Ciais, P.; Tømmervik, H.; et al. Characteristics, drivers and feedbacks of global greening. Nat. Rev. Earth Environ. 2019, 1, 14–27. [Google Scholar] [CrossRef]
  4. Fernández-Martínez, M.; Sardans, J.; Chevallier, F.; Ciais, P.; Obersteiner, M.; Vicca, S.; Canadell, J.G.; Bastos, A.; Friedlingstein, P.; Sitch, S.; et al. Global trends in carbon sinks and their relationships with CO2 and temperature. Nat. Clim. Chang. 2018, 9, 73–79. [Google Scholar] [CrossRef] [Green Version]
  5. WMO. WMO greenhouse gas bulletin. In The State of Greenhouse Gases in the Atmosphere Based on Global Observations Through 2020; WMO: Geneva, Switzerland, 2021. [Google Scholar]
  6. Lenton, T.M. Land and ocean carbon cycle feedback effects on global warming in a simple Earth system model. Tellus B Chem. Phys. Meteorol. 2000, 52, 1159–1188. [Google Scholar] [CrossRef] [Green Version]
  7. Sarmiento, J.L.; Gloor, M.; Gruber, N.; Beaulieu, C.; Jacobson, A.R.; Fletcher, S.E.M.; Pacala, S.; Rodgers, K. Trends and regional distributions of land and ocean carbon sinks. Biogeosciences 2010, 7, 2351–2367. [Google Scholar] [CrossRef] [Green Version]
  8. Kaushik, A.; Graham, J.; Dorheim, K.; Kramer, R.; Wang, J.; Byrne, B. The Future of the Carbon Cycle in a Changing Climate. Eos 2020, 101, 1728620. [Google Scholar] [CrossRef]
  9. Li, C.; Huang, J.; Ding, L.; Liu, X.; Han, D. Estimation of Oceanic and Land Carbon Sinks Based on the Most Recent Oxygen Budget. Earth’s Future 2021, 9, e2021EF002124. [Google Scholar] [CrossRef]
  10. Bousquet, P.; Peylin, P.; Ciais, P.; Le Quéré, C.; Friedlingstein, P.; Tans, P.P. Regional Changes in Carbon Dioxide Fluxes of Land and Oceans Since 1980. Science 2000, 290, 1342–1346. [Google Scholar] [CrossRef] [Green Version]
  11. Ciais, P.; Reichstein, M.; Viovy, N.; Granier, A.; Ogée, J.; Allard, V.; Aubinet, M.; Buchmann, N.; Bernhofer, C.; Carrara, A.; et al. Europe-wide reduction in primary productivity caused by the heat and drought in 2003. Nature 2005, 437, 529–533. [Google Scholar] [CrossRef]
  12. Goulden, M.L.; Munger, J.W.; Fan, S.-M.; Daube, B.C.; Wofsy, S.C. Exchange of Carbon Dioxide by a Deciduous Forest: Response to Interannual Climate Variability. Science 1996, 271, 1576–1578. [Google Scholar] [CrossRef] [Green Version]
  13. Wu, D.; Piao, S.; Zhu, D.; Wang, X.; Ciais, P.; Bastos, A.; Xu, X.; Xu, W. Accelerated terrestrial ecosystem carbon turnover and its drivers. Glob. Chang. Biol. 2020, 26, 5052–5062. [Google Scholar] [CrossRef] [PubMed]
  14. Spielmann, F.M.; Wohlfahrt, G.; Hammerle, A. Gross primary productivity of four European ecosystems constrained by joint CO2 and COS flux meas-urements. Geophys. Res. Lett. 2019, 46, 5284–5293. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Selmants, P.C.; Giardina, C.P.; Sousan, S.; Knapp, D.E.; Kimball, H.L.; Hawbaker, T.J.; Moreno, A.; Seirer, J.; Running, S.W.; Miura, T.; et al. Baseline carbon storage and carbon fluxes in terrestrial ecosystems of Hawai ‘i. In Baseline and Projected Future Carbon Storage and Carbon Fluxes in Ecosystems of Hawai’I; US Geological Survey Professional Paper 1834; US Department of the Interior, US Geological Survey: Reston, VA, USA, 2017; Chapter 6; pp. 75–87. [Google Scholar]
  16. Farquhar, G.D.; Von Caemmerer, S.; Berry, J.A. A biochemical model of photosynthetic CO2 assimilation in leaves of C3 species. Planta 1980, 149, 78–90. [Google Scholar] [CrossRef] [Green Version]
  17. Le Quéré, C.; Andrew, J.G.; Canadell, J.G.; Sitch, S.; Korsbakken, J.I.; Peters, G.P.; Manning, A.C.; Boden, T.A.; Tans, P.P.; Houghton, R.A.; et al. Global Carbon Budget 2016. Earth Syst. Sci. Data 2016, 8, 605–649. [Google Scholar] [CrossRef] [Green Version]
  18. Armond, P.A.; Schreiber, U.; Björkman, O. Photosynthetic Acclimation to Temperature in the Desert Shrub, Larrea divaricata: II. Light-harvesting Efficiency and Electron Transport 1. Plant Physiol. 1978, 61, 411–415. [Google Scholar] [CrossRef] [Green Version]
  19. Baldocchi, D.; Falge, E.; Gu, L.; Olson, R.; Hollinger, D.; Running, S.; Anthoni, P.; Bernhofer, C.; Davis, K.; Evans, R.; et al. FLUXNET: A New Tool to Study the Temporal and Spatial Variability of Ecosystem-Scale Carbon Dioxide, Water Vapor, and Energy Flux Densities. Bull. Am. Meteorol. Soc. 2001, 82, 2415–2434. [Google Scholar] [CrossRef]
  20. Niu, S.; Li, Z.; Xia, J.; Han, Y.; Wu, M.; Wan, S. Climatic warming changes plant photosynthesis and its temperature dependence in a temperate steppe of northern China. Environ. Exp. Bot. 2008, 63, 91–101. [Google Scholar] [CrossRef]
  21. Thornton, P.E.; Lamarque, J.-F.; Rosenbloom, N.A.; Mahowald, N. Influence of carbon-nitrogen cycle coupling on land model response to CO2fertilization and climate variability. Glob. Biogeochem. Cycles 2007, 21, GB4018. [Google Scholar] [CrossRef]
  22. Stevens, C.J.; David, T.I.; Storkey, J. Atmospheric nitrogen deposition in terrestrial ecosystems: Its impact on plant com-munities and consequences across trophic levels. Funct. Ecol. 2018, 32, 1757–1769. [Google Scholar] [CrossRef] [Green Version]
  23. Rustad, L.E. The response of terrestrial ecosystems to global climate change: Towards an integrated approach. Sci. Total Environ. 2008, 404, 222–235. [Google Scholar] [CrossRef]
  24. Nowak, R.S.; Ellsworth, D.S.; Smith, S.D. Functional responses of plants to elevated atmospheric CO2– do photosynthetic and productivity data from FACE experiments support early predictions? New Phytol. 2004, 162, 253–280. [Google Scholar] [CrossRef] [Green Version]
  25. McMurtrie, R.E.; Norby, R.; Medlyn, B.; Dewar, R.C.; Pepper, D.; Reich, P.; Barton, C.V.M. Why is plant-growth response to elevated CO2 amplified when water is limiting, but reduced when nitrogen is limiting? A growth-optimisation hypothesis. Funct. Plant Biol. 2008, 35, 521–534. [Google Scholar] [CrossRef] [PubMed]
  26. Chapin, F.S., III; Woodwell, G.M.; Randerson, J.T.; Rastetter, E.B.; Lovett, G.M.; Baldocchi, D.D.; Clark, D.A.; Harmon, M.E.; Schimel, D.S.; Valentini, R.; et al. Reconciling Carbon-cycle Concepts, Terminology, and Methods. Ecosystems 2006, 9, 1041–1050. [Google Scholar] [CrossRef] [Green Version]
  27. Lovett, G.M.; Cole, J.J.; Pace, M. Is Net Ecosystem Production Equal to Ecosystem Carbon Accumulation? Ecosystems 2006, 9, 152–155. [Google Scholar] [CrossRef]
  28. Xu, Q.; Hu, K.; Yao, Z.; Zuo, Q. Evaluation of carbon, nitrogen footprint and primary energy demand under different rice production systems. Ecol. Indic. 2020, 117, 106634. [Google Scholar] [CrossRef]
  29. Wagle, P.; Gowda, P.H.; Billesbach, D.P.; Northup, B.K.; Torn, M.S.; Neel, J.P.; Biraud, S.C. Dynamics of CO2 and H2O fluxes in Johnson grass in the U.S. Southern Great Plains. Sci. Total Environ. 2020, 739, 140077. [Google Scholar] [CrossRef]
  30. Campioli, M.; Malhi, Y.; Vicca, S.; Luyssaert, S.; Papale, D.; Peñuelas, J.; Reichstein, M.; Migliavacca, M.; Arain, M.A.; Janssens, I. Evaluating the convergence between eddy-covariance and biometric methods for assessing carbon budgets of forests. Nat. Commun. 2016, 7, 13717. [Google Scholar] [CrossRef]
  31. Baldocchi, D.D. Assessing the eddy covariance technique for evaluating carbon dioxide exchange rates of ecosystems: Past, present and future. Glob. Chang. Biol. 2003, 9, 479–492. [Google Scholar] [CrossRef] [Green Version]
  32. Kettle, A.J.; Kuhn, U.; von Hobe, M.; Kesselmeier, J.; Andreae, M. Global budget of atmospheric carbonyl sulfide: Temporal and spatial variations of the dominant sources and sinks. J. Geophys. Res. Earth Surf. 2002, 107, ACH 25-1–ACH 25-16. [Google Scholar] [CrossRef]
  33. Bingemer, H.G.; Bürgermeister, S.; Zimmermann, R.L.; Georgii, H.W. Atmospheric OCS: Evidence for a contribution of anthropogenic sources? J. Geophys. Res. 1990, 95(D12), 20617–20622. [Google Scholar] [CrossRef]
  34. Davidson, C.; Amrani, A.; Angert, A. Tropospheric carbonyl sulfide mass balance based on direct measurements of sulfur isotopes. Proc. Natl. Acad. Sci. USA 2021, 118. [Google Scholar] [CrossRef] [PubMed]
  35. Cutter, G.; Cutter, L.S.; Filippino, K.C. Sources and cycling of carbonyl sulfide in the Sargasso Sea. Limnol. Oceanogr. 2004, 49, 555–565. [Google Scholar] [CrossRef] [Green Version]
  36. Andreae, M. Ocean-atmosphere interactions in the global biogeochemical sulfur cycle. Mar. Chem. 1990, 30, 1–29. [Google Scholar] [CrossRef]
  37. Protoschill-Krebs, G.; Wilhelm, C.; Kesselmeier, J. Consumption of carbonyl sulphide (COS) by higher plant carbonic anhydrase (CA). Atmos. Environ. 1996, 30, 3151–3156. [Google Scholar] [CrossRef]
  38. Adams, D.F.; Farwell, S.O.; Pack, M.R.; Robinson, E. Biogenic Sulfur Gas Emissions from Soils in Eastern and Southeastern United States. J. Air Pollut. Control Assoc. 1981, 31, 1083–1089. [Google Scholar] [CrossRef]
  39. Lamb, B.; Westberg, H.; Allwine, G.; Bamesberger, L.; Guenther, A. Measurement of biogenic sulfur emissions from soils and vegetation: Application of dynamic enclosure methods with Natusch filter and GC/FPD analysis. J. Atmos. Chem. 1987, 5, 469–491. [Google Scholar] [CrossRef] [Green Version]
  40. Aneja, V.P.; Overton, J.H.; Cupitt, L.T., Jr.; Durham, J.L.; Wilson, W.E. Direct measurements of emission rates of some atmospheric biogenic sulfur compounds. Tellus 1979, 31, 174–178. [Google Scholar] [CrossRef] [Green Version]
  41. Goldan, P.D.; Fall, R.; Kuster, W.C.; Fehsenfeld, F.C. Uptake of COS by growing vegetation: A major tropospheric sink. J. Geophys. Res. Earth Surf. 1988, 93, 14186. [Google Scholar] [CrossRef]
  42. Whelan, M.E.; Min, D.-H.; Rhew, R.C. Salt marsh vegetation as a carbonyl sulfide (COS) source to the atmosphere. Atmospheric Environ. 2013, 73, 131–137. [Google Scholar] [CrossRef]
  43. Launois, T.; Belviso, S.; Bopp, L.; Fichot, C.G.; Peylin, P. A new model for the global biogeochemical cycle of carbonyl sulfide—Part 1: Assessment of direct marine emissions with an oceanic general circulation and biogeochemistry model. Atmospheric Chem. Phys. 2015, 15, 2295–2312. [Google Scholar] [CrossRef] [Green Version]
  44. Geng, C.; Mu, Y. Carbonyl sulfide and dimethyl sulfide exchange between lawn and the atmosphere. J. Geophys. Res. Atmos. 2004, 109, D12302. [Google Scholar] [CrossRef]
  45. Zumkehr, A.; Hilton, T.W.; Whelan, M.; Smith, S.; Campbell, J.E. Gridded anthropogenic emissions inventory and atmospheric transport of carbonyl sulfide in the U.S. J. Geophys. Res. Atmos. 2017, 122, 2169–2178. [Google Scholar] [CrossRef]
  46. Yujing, M.; Hai, W.; Zhang, X.; Jiang, G. Impact of anthropogenic sources on carbonyl sulfide in Beijing City. J. Geophys. Res. Earth Surf. 2002, 107, ACH 13-1–ACH 13-7. [Google Scholar] [CrossRef]
  47. Zumkehr, A.; Hilton, T.W.; Whelan, M.; Smith, S.; Kuai, L.; Worden, J.; Campbell, J.E. Global gridded anthropogenic emissions inventory of carbonyl sulfide. Atmos. Environ. 2018, 183, 11–19. [Google Scholar] [CrossRef]
  48. Khalil, M. Atmospheric trace gases, anthropogenic influences and global change. Encycl. Earth Syst. Sci. 1992, 1, 285–293. [Google Scholar]
  49. Lee, C.-L.; Brimblecombe, P. Anthropogenic contributions to global carbonyl sulfide, carbon disulfide and organosulfides fluxes. Earth-Sci. Rev. 2016, 160, 1–18. [Google Scholar] [CrossRef] [Green Version]
  50. Whelan, M.E.; Lennartz, S.T.; Gimeno, T.E.; Wehr, R.; Wohlfahrt, G.; Wang, Y.; Kooijmans, L.M.J.; Hilton, T.W.; Belviso, S.; Peylin, P.; et al. Reviews and syntheses: Carbonyl sulfide as a multi-scale tracer for carbon and water cycles. Biogeosciences 2018, 15, 3625–3657. [Google Scholar] [CrossRef] [Green Version]
  51. Stimler, K.; Montzka, S.; A Berry, J.; Rudich, Y.; Yakir, D. Relationships between carbonyl sulfide (COS) and CO 2 during leaf gas exchange. New Phytol. 2010, 186, 869–878. [Google Scholar] [CrossRef]
  52. Schenk, S.; Kesselmeier, J.; Anders, E. How does the exchange of one oxygen atom with sulfur affect the catalytic cycle of carbonic anhydrase? Chem. A Eur. J. 2004, 10, 3091–3105. [Google Scholar] [CrossRef]
  53. Notni, J.; Schenk, S.; Protoschill-Krebs, G.; Kesselmeier, J.; Anders, E. The missing link in COS metabolism: A model study on the reactivation of carbonic anhydrase from its hy-drosulfide analogue. ChemBioChem 2007, 8, 530–536. [Google Scholar] [CrossRef]
  54. Wohlfahrt, G.; Brilli, F.; Hörtnagl, L.; Xu, X.; Bingemer, H.; Hansel, A.; Loreto, F. Carbonyl sulfide (COS) as a tracer for canopy photosynthesis, transpiration and stomatal conductance: Potential and limitations†. Plant Cell Environ. 2011, 35, 657–667. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Simmons, J.S.; Klemedtsson, L.; Hultberg, H.; Hines, M.E. Consumption of atmospheric carbonyl sulfide by coniferous boreal forest soils. J. Geophys. Res. Earth Surf. 1999, 104, 11569–11576. [Google Scholar] [CrossRef]
  56. Seibt, U.; Kesselmeier, J.; Sandoval-Soto, L.; Kuhn, U.; Berry, J.A. A kinetic analysis of leaf uptake of COS and its relation to transpiration, photosynthesis and carbon isotope fractionation. Biogeosciences 2010, 7, 333–341. [Google Scholar] [CrossRef] [Green Version]
  57. Montzka, S.; Tans, P. Can Carbonyl Sulfide Help Constrain Gross Vegetative Fluxes of Carbon Dioxide? AGUFM 2004, 2004, B21E-04. [Google Scholar]
  58. Montzka, S.A.; Calvert, P.; Hall, B.D.; Elkins, J.W.; Conway, T.J.; Tans, P.P.; Sweeney, C. On the global distribution, seasonality, and budget of atmospheric carbonyl sulfide (COS) and some simi-larities to CO2. J. Geophys. Res. Atmos. 2007, 112, D09302. [Google Scholar] [CrossRef]
  59. Campbell, J.E.; Carmichael, G.R.; Chai, T.; Mena-Carrasco, M.; Tang, Y.; Blake, D.R.; Blake, N.J.; Vay, S.A.; Collatz, G.J.; Baker, I.; et al. Photosynthetic Control of Atmospheric Carbonyl Sulfide During the Growing Season. Science 2008, 322, 1085–1088. [Google Scholar] [CrossRef] [Green Version]
  60. Berry, J.; Wolf, A.; Campbell, J.E.; Baker, I.; Blake, N.; Blake, D.; Denning, A.S.; Kawa, S.R.; Montzka, S.; Seibt, U.; et al. A coupled model of the global cycles of carbonyl sulfide and CO2: A possible new window on the carbon cycle. J. Geophys. Res. Biogeosci. 2013, 118, 842–852. [Google Scholar] [CrossRef] [Green Version]
  61. Klein, T.; Rotenberg, E.; Cohen-Hilaleh, E.; Raz-Yaseef, N.; Tatarinov, F.; Preisler, Y.; Ogée, J.; Cohen, S.; Yakir, D. Quantifying transpirable soil water and its relations to tree water use dynamics in a water-limited pine forest. Ecohydrology 2013, 7, 409–419. [Google Scholar] [CrossRef]
  62. Oren, R.; Phillips, N.; Katul, G.; Ewers, B.; Pataki, D.E. Scaling xylem sap flux and soil water balance and calculating variance: A method for partitioning water flux in forests. EDP Sci. 1998, 55, 191–216. [Google Scholar] [CrossRef]
  63. Monteith, J.L.; Unsworth, M.H. Principles of environmental physics: Plants, Animals, and the Atmosphere, Fourth Edition; Academic Press: San Diego, CA, USA, 2013; pp. 181–271. ISBN 9780123869937. [Google Scholar]
  64. Gerdel, K.; Spielmann, F.M.; Hammerle, A.; Wohlfahrt, G. Eddy covariance carbonyl sulfide flux measurements with a quantum cascade laser absorption spectrometer. Atmos. Meas. Tech. 2017, 10, 3525–3537. [Google Scholar] [CrossRef] [Green Version]
  65. Kadmiel, M.; Berry, J.A.; Billesbach, D.; Campbell, J.E.; Torn, M.S.; Zahniser, M.; Seibt, U. Sources and sinks of carbonyl sulfide in an agricultural field in the Southern Great Plains. Proc. Natl. Acad. Sci. USA 2014, 111, 9064–9069. [Google Scholar]
  66. Kohonen, K.-M.; Kolari, P.; Kooijmans, L.M.J.; Chen, H.; Seibt, U.; Sun, W.; Mammarella, I. Towards standardized processing of eddy covariance flux measurements of carbonyl sulfide. Atmos. Meas. Tech. 2020, 13, 3957–3975. [Google Scholar] [CrossRef]
  67. Spielmann, F.M.; Hammerle, A.; Kitz, F.; Gerdel, K.; Wohlfahrt, G. Seasonal dynamics of the COS and CO2 exchange of a managed temperate grassland. Biogeosciences 2020, 17, 4281–4295. [Google Scholar] [CrossRef]
  68. Maignan, F.; Abadie, C.; Remaud, M.; Kooijmans, L.M.J.; Kohonen, K.-M.; Commane, R.; Wehr, R.; Campbell, J.E.; Belviso, S.; Montzka, S.A.; et al. Carbonyl sulfide: Comparing a mechanistic representation of the vegetation uptake in a land surface model and the leaf relative uptake approach. Biogeosci. 2021, 18, 2917–2955. [Google Scholar] [CrossRef]
  69. Powell, T.L.; Galbraith, D.R.; Christoffersen, B.O.; Harper, A.; Imbuzeiro, H.M.A.; Rowland, L.; Almeida, S.; Brando, P.M.; Costa, A.C.L.; Costa, M.H. Confronting model predictions of carbon fluxes with measurements of Amazon forests subjected to ex-perimental drought. New Phytol. 2013, 200, 350–365. [Google Scholar] [CrossRef] [Green Version]
  70. Kooijmans, L.M.J.; Kooijmans, L.M.J.; Cho, A.; Ma, J.; Kaushik, A.; Haynes, K.D.; Baker, I.; Ingrid, T.; Luijkx, I.T.; Groenink, M.; et al. Evaluation of carbonyl sulfide biosphere exchange in the Simple Biosphere Model (SiB4). Biogeosci. Discuss. 2021, 2021, 1–32. [Google Scholar] [CrossRef]
  71. Hu, L.; Montzka, S.A.; Kaushik, A.; Andrews, A.E.; Sweeney, C.; Miller, J.; Baker, J.T.; Denning, S.; Campbell, E.; Shiga, Y.P. COS-derived GPP relationships with temperature and light help explain high-latitude atmospheric CO2 seasonal cycle amplification. Proc. Natl. Acad. Sci. USA 2021, 118, e2103423118. [Google Scholar] [CrossRef]
  72. Cochavi, A.; Amer, M.; Stern, R.; Tatarinov, F.; Migliavacca, M.; Yakir, D. Differential responses to two heatwave intensities in a Mediterranean citrus orchard are identified by combining measurements of fluorescence and carbonyl sulfide (COS) and CO2 uptake. New Phytol. 2021, 230, 1394–1406. [Google Scholar] [CrossRef]
  73. Wehr, R.; Commane, R.; Munger, J.W.; McManus, J.B.; Nelson, D.D.; Zahniser, M.S.; Saleska, S.R.; Wofsy, S.C. Dynamics of canopy stomatal conductance, transpiration, and evaporation in a temperate deciduous forest, validated by carbonyl sulfide uptake. Biogeosciences 2017, 14, 389–401. [Google Scholar] [CrossRef] [Green Version]
  74. Zhang, Y.; Bastos, A.; Maignan, F.; Goll, D.; Boucher, O.; Li, L.; Cescatti, A.; Vuichard, N.; Chen, X.; Ammann, C. Modeling the impacts of diffuse light fraction on photosynthesis in ORCHIDEE (v5453) land surface model. Geosci. Model Dev. 2020, 13, 5401–5423. [Google Scholar] [CrossRef]
  75. Wehr, R.; Saleska, S.R. Calculating canopy stomatal conductance from eddy covariance measurements, in light of the energy budget closure problem. Biogeosciences 2021, 18, 13–24. [Google Scholar] [CrossRef]
  76. Camy-Peyret, C.; Liuzzi, G.; Masiello, G.; Serio, C.; Venafra, S.; Montzka, S. Assessment of IASI capability for retrieving carbonyl sulphide (OCS). J. Quant. Spectrosc. Radiat. Transf. 2017, 201, 197–208. [Google Scholar] [CrossRef]
  77. Serio, C.; Masiello, G.; Mastro, P.; Belviso, S.; Remaud, M. Seasonal variability of degrees of freedom and its effect over time series and spatial patterns of atmospheric gases from satellite: Application to carbonyl sulfide (OCS). In Proceedings of the Remote Sensing of Clouds and the Atmosphere XXVI, Online Only, 12 September 2021. [Google Scholar] [CrossRef]
  78. Barkley, M.P.; Palmer, P.I.; Boone, C.D.; Bernath, P.F.; Suntharalingam, P. Global distributions of carbonyl sulfide in the upper troposphere and stratosphere. Geophys. Res. Lett. 2008, 35. [Google Scholar] [CrossRef] [Green Version]
  79. Glatthor, N.; Höpfner, M.; Leyser, A.; Stiller, G.P.; von Clarmann, T.; Grabowski, U.; Kellmann, S.; Linden, A.; Sinnhuber, B.-M.; Krysztofiak, G.; et al. Global carbonyl sulfide (OCS) measured by MIPAS/Envisat during 2002–2012. Atmos. Chem. Phys. 2017, 17, 2631–2652. [Google Scholar] [CrossRef] [Green Version]
  80. Yousefi, M.; Bernatha, P.F.; Boonec, C.D.; Toond, G.C. Global measurements of atmospheric carbonyl sulfide (OCS), OC34S and O13CS. J. Quant. Spec-Troscopy Radiat. Transf. 2019, 238, 106554. [Google Scholar] [CrossRef]
  81. Knohl, A.; Cuntz, M. Tracing carbon fixation. Nat. Clim. Chang. 2017, 7, 393–394. [Google Scholar] [CrossRef]
  82. Stinecipher, J.; Cameron-Smith, P.; Blake, N.; Kuai, L.; Lejeune, B.; Mahieu, E.; Simpson, I.; Campbell, J. Biomass Burning Unlikely to Account for Missing Source of Carbonyl Sulfide. Geophys. Res. Lett. 2019, 46, 14912–14920. [Google Scholar] [CrossRef]
  83. Campbell, J.E.; Whelan, M.E.; Seibt, U.; Smith, S.J.; Berry, J.A.; Hilton, T.W. Atmospheric carbonyl sulfide sources from anthropogenic activity: Implications for carbon cycle con-straints. Geophys. Res. Lett. 2015, 42, 3004–3010. [Google Scholar] [CrossRef]
  84. Hattori, S.; Kamezaki, K.; Yoshida, N. Constraining the atmospheric OCS budget from sulfur isotopes. Proc. Natl. Acad. Sci. USA 2020, 117, 20447–20452. [Google Scholar] [CrossRef]
  85. Gimeno, T.E.; Ogee, J.; Royles, J.; Gibon, Y.; West, J.B.; Burlett, R.; Jones, S.; Sauze, J.; Wohl, S.; Benard, C.; et al. Bryophyte gas-exchange dynamics along varying hydration status reveal a significant carbonyl sulphide (COS) sink in the dark and COS source in the light. New Phytol. 2017, 215, 965–976. [Google Scholar] [CrossRef] [Green Version]
  86. Ogawa, T.; Noguchi, K.; Saito, M.; Nagahata, Y.; Kato, H.; Ohtaki, A.; Nakayama, H.; Dohmae, N.; Matsushita, Y.; Odaka, M.; et al. Carbonyl Sulfide Hydrolase from Thiobacillus thioparus Strain THI115 Is One of the β-Carbonic Anhydrase Family Enzymes. J. Am. Chem. Soc. 2013, 135, 3818–3825. [Google Scholar] [CrossRef] [PubMed]
  87. Lorimer, G.H.; Pierce, J. Carbonyl sulfide: An alternate substrate for but not an activator of ribulose-1,5-bisphosphate carboxylase. J. Biol. Chem. 1989, 264, 2764–2772. [Google Scholar] [CrossRef]
  88. Ogée, J.; Sauze, J.; Kesselmeier, J.; Genty, B.; Van Diest, H.; Launois, T.; Wingate, L. A new mechanistic framework to predict OCS fluxes from soils. Biogeosciences 2016, 13, 2221–2240. [Google Scholar] [CrossRef] [Green Version]
  89. Tcherkez, G.G.B.; Farquhar, G.; Andrews, T.J. Despite slow catalysis and confused substrate specificity, all ribulose bisphosphate carboxylases may be nearly perfectly optimized. Proc. Natl. Acad. Sci. USA 2006, 103, 7246–7251. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Yang, F.; Qubaja, R.; Tatarinov, F.; Rotenberg, E.; Yakir, D. Assessing canopy performance using carbonyl sulfide measurements. Glob. Chang. Biol. 2018, 24, 3486–3498. [Google Scholar] [CrossRef] [PubMed]
  91. Nadezhdina, N.; Plichta, R.; Nadezhdin, V.; Gebauer, R.; Jupa, R.; Habrova, H.; Madera, P. A comparative structural and functional study of leaf traits and sap flow in Dracaena cinnabari and Dracaena draco seedlings. Funct. Plant Biol. 2015, 42, 1092. [Google Scholar] [CrossRef]
  92. Leakey, A.D.B.; Ferguson, J.N.; Pignon, C.P.; Wu, A.; Jin, Z.; Hammer, G.L.; Lobell, D.B. Water Use Efficiency as a Constraint and Target for Improving the Resilience and Productivity of C-3 and C-4 Crops. In Annual Review of Plant Biology; Merchant, S.S., Ed.; Annual Reviews: Palo Alto, CA, USA, 2019; Volume 70, pp. 781–808. [Google Scholar]
  93. Gago, J.; Douthe, C.; Florez-Sarasa, I.; Escalona, J.M.; Galmés, J.; Fernie, A.R.; Flexas, J.; Medrano, H. Opportunities for improving leaf water use efficiency under climate change conditions. Plant Sci. 2014, 226, 108–119. [Google Scholar] [CrossRef]
  94. Gago, J.; Daloso, D.; Carriquí, M.; Nadal, M.; Morales, M.; Araújo, W.; Nunes-Nesi, A.; Perera-Castro, A.; Clemente-Moreno, M.J.; Flexas, J. The photosynthesis game is in the "inter-play": Mechanisms underlying CO2 diffusion in leaves. Environ. Exp. Bot. 2020, 178, 104174. [Google Scholar] [CrossRef]
  95. Warren, J.M.; Pötzelsberger, E.; Wullschleger, S.; Thornton, P.; Hasenauer, H.; Norby, R. Ecohydrologic impact of reduced stomatal conductance in forests exposed to elevated CO2. Ecohydrology 2010, 4, 196–210. [Google Scholar] [CrossRef]
  96. Sandoval-Soto, L.; Stanimirov, M.; von Hobe, M.; Schmitt, V.; Valdes, J.; Wild, A.; Kesselmeier, J. Global uptake of carbonyl sulfide (COS) by terrestrial vegetation: Estimates corrected by deposition velocities normalized to the uptake of carbon dioxide (CO2). Biogeosciences 2005, 2, 125–132. [Google Scholar] [CrossRef] [Green Version]
  97. Gillon, J.S.; Yakir, D. Internal Conductance to CO2 Diffusion and C18OO Discrimination in C3 Leaves. Plant Physiol. 2000, 123, 201–214. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  98. Frank, D.C.; Poulter, B.; Saurer, M.; Esper, J.; Huntingford, C.; Helle, G.; Treydte, K.; Zimmermann, N.E.; Schleser, G.H.; Ahlström, A.; et al. Water-use efficiency and transpiration across European forests during the Anthropocene. Nat. Clim. Chang. 2015, 5, 579–583. [Google Scholar] [CrossRef] [Green Version]
  99. Yonemura, S.; Sandoval-Soto, L.; Kesselmeier, J.; Kuhn, U.; Hobe, M.W.; Yakir, D.; Kawashima, S. Uptake of carbonyl sulfide (COS) and emission of dimethyl sulfide (DMS) by plants. In Proceedings of the 6th International Symposium on Plant Responses to Air Pollution and Global Changes, Tsukuba, Japan, 19–22 October 2004. [Google Scholar]
  100. Gai, C.; Dong, Y.; Zhang, T. Distribution of sulfur species in gaseous and condensed phase during downdraft gasification of corn straw. Energy. 2014, 64, 248–258. [Google Scholar] [CrossRef]
  101. Commane, R.; Meredith, L.; Baker, I.T.; Berry, J.A.; Munger, J.W.; Montzka, S.; Templer, P.H.; Juice, S.; Zahniser, M.S.; Wofsy, S.C. Seasonal fluxes of carbonyl sulfide in a midlatitude forest. Proc. Natl. Acad. Sci. USA 2015, 112, 14162–14167. [Google Scholar] [CrossRef] [Green Version]
  102. Ball, J.T. An Analysis of Stomatal Conductance; Stanford University Stanford: Stanford, CA, USA, 1988. [Google Scholar]
  103. Collatz, G.; Ball, J.; Grivet, C.; A Berry, J. 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]
  104. Leuning, R. A critical appraisal of a combined stomatal-photosynthesis model for C3 plants. Plant Cell Environ. 1995, 18, 339–355. [Google Scholar] [CrossRef]
  105. Farquhar, G.D.; Sharkey, T. Stomatal Conductance and Photosynthesis. Annu. Rev. Plant Physiol. 1982, 33, 317–345. [Google Scholar] [CrossRef]
  106. Liu, N.; Kala, J.; Liu, S.; Haverd, V.; Dell, B.; Smettem, K.R.; Harper, R.J. Drought can offset potential water use efficiency of forest ecosystems from rising atmospheric CO2. J. Environ. Sci. 2019, 90, 262–274. [Google Scholar] [CrossRef] [PubMed]
  107. Umair, M.; Kim, D.; Choi, M. Impact of climate, rising atmospheric carbon dioxide, and other environmental factors on water-use efficiency at multiple land cover types. Sci. Rep. 2020, 10, 1–13. [Google Scholar] [CrossRef]
  108. Xu, H.; Zhang, Z.; Xiao, J.; Chen, J.; Zhu, M.; Cao, W.; Chen, Z. Environmental and canopy stomatal control on ecosystem water use efficiency in a riparian poplar plantation. Agric. For. Meteorol. 2020, 287, 107953. [Google Scholar] [CrossRef]
  109. Wingate, L.; Ogée, J.; Cuntz, M.; Genty, B.; Reiter, I.; Seibt, U.; Yakir, D.; Maseyk, K.; Pendall, E.G.; Barbour, M.M. The impact of soil microorganisms on the global budget of δ18O in atmospheric CO2. Proc. Natl. Acad. Sci. USA 2009, 106, 22411. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  110. De Kauwe, M.G.; Medlyn, B.; Zaehle, S.; Walker, A.P.; Dietze, M.; Hickler, T.; Jain, A.K.; Luo, Y.; Parton, W.J.; Prentice, I.C.; et al. Forest water use and water use efficiency at elevated CO2: A model-data intercomparison at two contrasting temperate forest FACE sites. Glob. Chang. Biol. 2013, 19, 1759–1779. [Google Scholar] [CrossRef] [PubMed]
  111. Reichstein, M.; Falge, E.; Baldocchi, D.; Papale, D.; Aubinet, M.; Berbigier, P.; Bernhofer, C.; Buchmann, N.; Gilmanov, T.; Granier, A.; et al. On the separation of net ecosystem exchange into assimilation and ecosystem respiration: Review and improved algorithm. Glob. Chang. Biol. 2005, 11, 1424–1439. [Google Scholar] [CrossRef]
  112. Sandoval-Soto, L.; Kesselmeier, M.; Schmitt, V.; Wild, A.; Kesselmeier, J. Observations of the uptake of carbonyl sulfide (COS) by trees under elevated atmospheric carbon dioxide concentrations. Biogeosciences Discuss. 2012, 9, 2123–2152. [Google Scholar] [CrossRef] [Green Version]
  113. Berkelhammer, M.; Asaf, D.; Still, C.; Montzka, S.; Noone, D.; Gupta, M.; Provencal, R.; Chen, H.; Yakir, D. Constraining surface carbon fluxes using in situ measurements of carbonyl sulfide and carbon dioxide. Glob. Biogeochem. Cycles 2014, 28, 161–179. [Google Scholar] [CrossRef] [Green Version]
  114. Billesbach, D.P.; Berry, J.A.; Seibt, U.; Maseyk, K.; Torn, M.S.; Fischer, M.L.; Abu-Naser, M.; Campbell, J.E. Growing season eddy covariance measurements of carbonyl sulfide and CO2 fluxes: COS and CO2 rela-tionships in Southern Great Plains winter wheat. Agric. For. Meteorol. 2014, 184, 48–55. [Google Scholar] [CrossRef]
  115. Kooijmans, L.M.J.; Sun, W.; Aalto, J.; Erkkilä, K.-M.; Maseyk, K.; Seibt, U.; Vesala, T.; Mammarella, I.; Chen, H. Influences of light and humidity on carbonyl sulfide-based estimates of photosynthesis. Proc. Natl. Acad. Sci. USA 2019, 116, 2470–2475. [Google Scholar] [CrossRef] [Green Version]
  116. Blonquist, J.M.; Montzka, S.A.; Munger, J.W.; Yakir, D.; Desai, A.R.; Dragoni, D.; Griffis, T.J.; Monson, R.K.; Scott, R.L.; Bowling, D.R. The potential of carbonyl sulfide as a proxy for gross primary production at flux tower sites. J. Geophys. Res. Biogeosci. 2011, 116, G04019. [Google Scholar] [CrossRef] [Green Version]
  117. Stimler, K.; A Berry, J.; Montzka, S.A.; Yakir, D. Association between Carbonyl Sulfide Uptake and 18Δ during Gas Exchange in C3 and C4 Leaves. Plant Physiol. 2011, 157, 509–517. [Google Scholar] [CrossRef] [Green Version]
  118. Kooijmans, L.M.J.; Maseyk, K.; Seibt, U.; Sun, W.; Vesala, T.; Mammarella, I.; Kolari, P.; Aalto, J.; Franchin, A.; Vecchi, R.; et al. Canopy uptake dominates nighttime carbonyl sulfide fluxes in a boreal forest. Atmos. Chem. Phys. 2017, 17, 11453–11465. [Google Scholar] [CrossRef] [Green Version]
  119. Jing, W.; Wang, L.; Li, D.; Bao, X.; Shi, Y. Carbonyl sulfide (COS) and carbon disulfide (CS2) exchange fluxes between cotton fields and the atmosphere in the arid area in Xinjiang, China. Environ. Geochem. Health 2019, 41, 2195–2207. [Google Scholar] [CrossRef] [PubMed]
  120. Bartell, U.; Hofmann, U.; Hofmann, R.; Kreuzburg, B.; Andreae, M.; Kesselmeier, J. COS and H2S fluxes over a wet meadow in relation to photosynthetic activity: An analysis of measurements made on 6 September 1990. Atmos. Environ. Part A Gen. Top. 1993, 27, 1851–1864. [Google Scholar] [CrossRef] [Green Version]
  121. Ewers, B.E.; Oren, R.; Kim, H.S.; Bohrer, G.; Lai, C.T. Effects of hydraulic architecture and spatial variation in light on mean stomatal conductance of tree branches and crowns. Plant Cell Environ. 2007, 30, 483–496. [Google Scholar] [CrossRef] [PubMed]
  122. Asaf, D.; Rotenberg, E.; Tatarinov, F.; Dicken, U.; Montzka, S.; Yakir, D. Ecosystem photosynthesis inferred from measurements of carbonyl sulphide flux. Nat. Geosci. 2013, 6, 186–190. [Google Scholar] [CrossRef]
  123. Rastogi, B.; Berkelhammer, M.; Wharton, S.; Whelan, M.E.; Meinzer, F.C.; Noone, D.; Still, C.J. Ecosystem fluxes of carbonyl sulfide in an old-growth forest: Temporal dynamics and responses to diffuse radiation and heat waves. Biogeosciences 2018, 15, 7127–7139. [Google Scholar] [CrossRef] [Green Version]
  124. Stimler, K.; A Berry, J.; Yakir, D. Effects of Carbonyl Sulfide and Carbonic Anhydrase on Stomatal Conductance. Plant Physiol. 2011, 158, 524–530. [Google Scholar] [CrossRef]
  125. Rastogi, B.; Berkelhammer, M.; Wharton, S.; Whelan, M.E.; Itter, M.S.; Leen, J.B.; Gupta, M.X.; Noone, D.; Still, C.J. Large Uptake of Atmospheric OCS Observed at a Moist Old Growth Forest: Controls and Implications for Carbon Cycle Applications. J. Geophys. Res. Biogeosci. 2018, 123, 3424–3438. [Google Scholar] [CrossRef]
  126. Berkelhammer, M.; Alsip, B.; Matamala, R.; Cook, D.; Whelan, M.E.; Joo, E.; Bernacchi, C.; Miller, J.; Meyers, T. Seasonal Evolution of Canopy Stomatal Conductance for a Prairie and Maize Field in the Midwestern United States from Continuous Carbonyl Sulfide Fluxes. Geophys. Res. Lett. 2020, 47. [Google Scholar] [CrossRef]
  127. Sun, W.; Maseyk, K.; Lett, C.; Seibt, U. Stomatal control of leaf fluxes of carbonyl sulfide and CO2 in a Typha freshwater marsh. Biogeosciences 2018, 15, 3277–3291. [Google Scholar] [CrossRef] [Green Version]
  128. Xu, X.; Bingemer, H.G.; Schmidt, U. The flux of carbonyl sulfide and carbon disulfide between the atmosphere and a spruce forest. Atmos. Chem. Phys. Discuss. 2002, 2, 171–181. [Google Scholar] [CrossRef] [Green Version]
  129. White, M.L.; Zhou, Y.; Russo, R.S.; Mao, H.; Talbot, R.; Varner, R.K.; Sive, B.C. Carbonyl sulfide exchange in a temperate loblolly pine forest grown under ambient and elevated CO2. Atmos. Chem. Phys. 2010, 9, 17219–17263. [Google Scholar]
  130. MacDowall, F.D.H. Midday closure of stomata in aging tobacco leaves. Can. J. Bot. 1963, 41, 1289–1300. [Google Scholar] [CrossRef]
  131. Rosenberg, M.S. MetaWin: Statistical Software for Meta-Analysis: Version 2; Sinauer: Sunderland, MA, USA, 2000. [Google Scholar]
  132. Hedges, L.V.; Gurevitch, J.; Curtis, P.S. The Meta-Analysis of Response Ratios in Experimental Ecology. Ecology 1999, 80, 1150–1156. [Google Scholar] [CrossRef]
  133. De Pury, D.G.G.; Farquhar, G.D. Simple scaling of photosynthesis from leaves to canopies without the errors of big-leaf models. Plant Cell Environ. 1997, 20, 537–557. [Google Scholar] [CrossRef]
  134. Pollok, M.; Heber, U.; Naik, M.S. Inhibition of stomatal opening in sunflower leaves by car-bon-monoxide, and reversal of inhibition by light. Planta 1989, 178, 223–230. [Google Scholar] [CrossRef]
  135. Costa, J.; Monnet, F.; Jannaud, D.; Leonhardt, N.; Ksas, B.; Reiter, I.; Pantin, F.; Genty, B. OPEN ALL NIGHT LONG: The Dark Side of Stomatal Control. Plant Physiol. 2014, 167, 289–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Bloem, E.; Haneklaus, S.; Kesselmeier, J.; Schnug, E. Sulfur Fertilization and Fungal Infections Affect the Exchange of H2S and COS from Agricultural Crops. J. Agric. Food Chem. 2012, 60, 7588–7596. [Google Scholar] [CrossRef] [PubMed]
  137. Wagner, S.; Zotz, G.; Allen, N.S.; Bader, M.Y. Altitudinal changes in temperature responses of net photosynthesis and dark respiration in tropical bryo-phytes. Ann. Bot. 2013, 111, 455–465. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  138. Smith, E.C.; Griffiths, H. The role of carbonic anhydrase in photosynthesis and the activity of the carbon-concentrating-mechanism in bryophytes of the class Anthocerotae. New Phytol. 2000, 145, 29–37. [Google Scholar] [CrossRef]
  139. Badger, M. The roles of carbonic anhydrases in photosynthetic CO2 concentrating mechanisms. Photosynth. Res. 2003, 77, 83–94. [Google Scholar] [CrossRef]
  140. Meyer, M.; Seibt, U.; Griffiths, H. To concentrate or ventilate? Carbon acquisition, isotope discrimination and physiological ecology of early land plant life forms. Philos. Trans. R. Soc. B Biol. Sci. 2008, 363, 2767–2778. [Google Scholar] [CrossRef] [Green Version]
  141. Watts, S.f. The mass budgets of carbonyl sulfide, dimethyl sulfide, carbon disulfide and hydrogen sulfide. Atmos. Environ. 2000, 34, 761–779. [Google Scholar] [CrossRef]
  142. Masaki, Y.; Ozawa, R.; Kageyama, K.; Katayama, Y. Degradation and emission of carbonyl sulfide, an atmospheric trace gas, by fungi isolated from forest soil. FEMS Microbiol. Lett. 2016, 363, fnw197. [Google Scholar] [CrossRef] [PubMed]
  143. Ogawa, T.; Kato, H.; Higashide, M.; Nishimiya, M.; Katayama, Y. Degradation of carbonyl sulfide by Actinomycetes and detection of clade D of β-class carbonic anhydrase. FEMS Microbiol. Lett. 2016, 363, fnw223. [Google Scholar] [CrossRef] [PubMed]
  144. Kesselmeier, J.; Merk, L. Exchange of carbonyl sulfide (COS) between agricultural plants and the atmosphere: Studies on the deposition of COS to peas, corn and rapeseed. Biogeochemistry 1993, 23, 47–59. [Google Scholar] [CrossRef]
  145. Majumdar, S.; Ghosh, S.; Glick, B.R.; Dumbroff, E.B. Activities of chlorophyllase, phosphoenolpyruvate carboxylase and ribulose-1, 5-bisphosphate carboxylase in the primary leaves of soybean during senescence and drought. Physiol. Plant. 1991, 81, 473–480. [Google Scholar] [CrossRef]
  146. Khanna-Chopra, R. Leaf senescence and abiotic stresses share reactive oxygen species-mediated chloroplast degradation. Protoplasma 2011, 249, 469–481. [Google Scholar] [CrossRef] [PubMed]
  147. Zabaleta, E.; Martin, M.V.; Braun, H.-P. A basal carbon concentrating mechanism in plants? Plant Sci. 2012, 187, 97–104. [Google Scholar] [CrossRef]
  148. Rech, M.; Morant-Manceau, A.; Tremblin, G. Carbon fixation and carbonic anhydrase activity in Haslea ostrearia (Bacil-lariophyceae) in relation to growth irradiance. Photosynthetica 2008, 46, 56–62. [Google Scholar] [CrossRef]
  149. Geng, C.; Mu, Y. Carbonyl sulfide and dimethyl sulfide exchange between trees and the atmosphere. Atmos. Environ. 2006, 40, 1373–1383. [Google Scholar] [CrossRef]
  150. Biscoe, P.V.; Littleton, E.J.; Scott, R.K. Stomatal control of gas exchange in barley awns. Ann. Appl. Biol. 1973, 75, 285–297. [Google Scholar] [CrossRef]
  151. Lloyd, J.; Farquhar, G. Effects of rising temperatures and [CO 2 ] on the physiology of tropical forest trees. Philos. Trans. R. Soc. B Biol. Sci. 2008, 363, 1811–1817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  152. Wehr, R.; Munger, J.W.; McManus, J.B.; Nelson, D.D.; Zahniser, M.S.; Davidson, E.A.; Wofsy, S.C.; Saleska, S.R. Seasonality of temperate forest photosynthesis and daytime respiration. Nature 2016, 534, 680–683. [Google Scholar] [CrossRef]
  153. Caird, M.A.; Richards, J.H.; Donovan, L. Nighttime Stomatal Conductance and Transpiration in C3 and C4 Plants. Plant Physiol. 2007, 143, 4–10. [Google Scholar] [CrossRef] [Green Version]
  154. Belviso, S.; Schmidt, M.; Yver, C.; Ramonet, M.; Gros, V.; Launois, T. Strong similarities between night-time deposition velocities of carbonyl sulphide and molecular hydrogen inferred from semi-continuous atmospheric observations in Gif-sur-Yvette, Paris region. Tellus B Chem. Phys. Meteorol. 2013, 65. [Google Scholar] [CrossRef] [Green Version]
  155. Commane, R.; Herndon, S.C.; Zahniser, M.S.; Lerner, B.M.; McManus, J.B.; Munger, J.W.; Nelson, D.D.; Wofsy, S.C. Carbonyl sulfide in the planetary boundary layer: Coastal and continental influences. J. Geophys. Res. Atmos. 2013, 118, 8001–8009. [Google Scholar] [CrossRef] [Green Version]
  156. Whelan, M.E.; Hilton, T.W.; Berry, J.A.; Berkelhammer, M.; Desai, A.R.; Campbell, J.E. Carbonyl sulfide exchange in soils for better estimates of ecosystem carbon uptake. Atmos. Chem. Phys. 2016, 16, 3711–3726. [Google Scholar] [CrossRef] [Green Version]
  157. Kuhn, U. Carbonyl sulfide exchange on an ecosystem scale: Soil represents a dominant sink for atmospheric COS. Atmos. Environ. 1999, 33, 995–1008. [Google Scholar] [CrossRef]
  158. Kesselmeier, J.; Meixner, F.X.; Hofmann, U.; Ajavon, A.L.; Leimbach, S.; Andreae, M.O. Reduced sulfur compound exchange between the atmosphere and tropical tree species in southern Cameroon. Biogeochemistry 1993, 23, 23–45. [Google Scholar] [CrossRef]
  159. Porada, P.; Weber, B.; Elbert, W.; Pöschl, U.; Kleidon, A. Estimating global carbon uptake by lichens and bryophytes with a process-based model. Biogeosciences 2013, 10, 6989–7033. [Google Scholar] [CrossRef] [Green Version]
Figure 2. Locations of carbonyl sulfide uptake observation sites in this meta-analysis study. The full name of each biome, as well as its source literature are NLF—needle leaf forest [113,115,118,123,125,128,129], BLF—broad leaf forest [14,73,90,101,113,129], MXF—mixed forest [73,96,115], GRA—grassland [14,113,114,120,126], MSH—marshland [127], CPLC3—cropland dominated by C3 species [14,51,65,114,117,119,122,124] and CPLC4—cropland dominated by C4 species [117,124,126].
Figure 2. Locations of carbonyl sulfide uptake observation sites in this meta-analysis study. The full name of each biome, as well as its source literature are NLF—needle leaf forest [113,115,118,123,125,128,129], BLF—broad leaf forest [14,73,90,101,113,129], MXF—mixed forest [73,96,115], GRA—grassland [14,113,114,120,126], MSH—marshland [127], CPLC3—cropland dominated by C3 species [14,51,65,114,117,119,122,124] and CPLC4—cropland dominated by C4 species [117,124,126].
Sustainability 14 02840 g002
Figure 3. Environmental condition changes between the control and the treatment groups in each study. Boxes (dots) and the error bars represent the average changes in environmental conditions and the extent of these changes. The number next to the dots is the sample size for each variable. (a) LI, the light intensity (μmol m−2 s−1); (b) temperature, the atmospheric temperature (°C); (c) rH, the relative humidity (%); (d) VPD, the vapor pressure deficit (Pa); (e) Ca,CO2, the atmospheric CO2 concentration (ppt); (f) Ca,COS, the atmospheric COS concentration (ppm). NVS.ME, the noon period (9:00~15:00) in comparison with the periods of morning (6:00~9:00) and evening (15:00~18:00); FAVS.DD, the forenoon (8:00~11:00) and afternoon (15:00~17:00) periods in comparison with the periods of dawn (5:00~8:00) and dusk (17:00~20:00); D vs. N, the daytime (5:00~20:00) in comparison with nighttime (20:00~5:00); daily, the daily observations separated into the treatment (sunny) and the control (cloudy) groups by light intensity (PAR = 600 μmolm−2s−1); seasonal, the growing season continuous observations separated into the treatment (summer) and the control (spring and autumn) groups by the vernal equinox day (DOY 171) and the autumnal equinox day (DOY 266); control, the light control experiment separated into the treatment and the control groups by the standard of light intensity (PAR = 600 μmol m−2 s−1).
Figure 3. Environmental condition changes between the control and the treatment groups in each study. Boxes (dots) and the error bars represent the average changes in environmental conditions and the extent of these changes. The number next to the dots is the sample size for each variable. (a) LI, the light intensity (μmol m−2 s−1); (b) temperature, the atmospheric temperature (°C); (c) rH, the relative humidity (%); (d) VPD, the vapor pressure deficit (Pa); (e) Ca,CO2, the atmospheric CO2 concentration (ppt); (f) Ca,COS, the atmospheric COS concentration (ppm). NVS.ME, the noon period (9:00~15:00) in comparison with the periods of morning (6:00~9:00) and evening (15:00~18:00); FAVS.DD, the forenoon (8:00~11:00) and afternoon (15:00~17:00) periods in comparison with the periods of dawn (5:00~8:00) and dusk (17:00~20:00); D vs. N, the daytime (5:00~20:00) in comparison with nighttime (20:00~5:00); daily, the daily observations separated into the treatment (sunny) and the control (cloudy) groups by light intensity (PAR = 600 μmolm−2s−1); seasonal, the growing season continuous observations separated into the treatment (summer) and the control (spring and autumn) groups by the vernal equinox day (DOY 171) and the autumnal equinox day (DOY 266); control, the light control experiment separated into the treatment and the control groups by the standard of light intensity (PAR = 600 μmol m−2 s−1).
Sustainability 14 02840 g003aSustainability 14 02840 g003b
Figure 4. The averages of Hedges d and LnRR , with its’ 95% confidence intervals (CI) of the dependent variables in each group. In the subfigures, dependent variables are (a) LRU, (b) gs, (c) FC,COS, (d) FC,CO2, separately. The control and and treatment samples are shown in Tables S2 and S3. If the 95% CI does not include 0, the dependent variables were significantly affected by the change in environmental conditions between the control and the treatment. The dashed vertical line indicates Hedges’d (LnRR) = 0. The number next to the dots is the sample size for each variable.
Figure 4. The averages of Hedges d and LnRR , with its’ 95% confidence intervals (CI) of the dependent variables in each group. In the subfigures, dependent variables are (a) LRU, (b) gs, (c) FC,COS, (d) FC,CO2, separately. The control and and treatment samples are shown in Tables S2 and S3. If the 95% CI does not include 0, the dependent variables were significantly affected by the change in environmental conditions between the control and the treatment. The dashed vertical line indicates Hedges’d (LnRR) = 0. The number next to the dots is the sample size for each variable.
Sustainability 14 02840 g004aSustainability 14 02840 g004b
Figure 5. Relationship of the Hedges   d of the LRU with the variation in environmental conditions in the diurnal cycling groups. Subfingures (af) denote the linear regression between the Hedges d value of LRU and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Figure 5. Relationship of the Hedges   d of the LRU with the variation in environmental conditions in the diurnal cycling groups. Subfingures (af) denote the linear regression between the Hedges d value of LRU and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Sustainability 14 02840 g005
Figure 6. Relationship of the Hedges   d of the gs with the variation in environmental conditions in the diurnal cycling groups. Subfingures (af) denote the linear regression between the Hedges d value of gs and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Figure 6. Relationship of the Hedges   d of the gs with the variation in environmental conditions in the diurnal cycling groups. Subfingures (af) denote the linear regression between the Hedges d value of gs and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Sustainability 14 02840 g006
Figure 7. Relationship of the Hedges   d of FC,COS with the variation in environmental conditions in the diurnal cycling groups. Subfingures (af) denote the linear regression between the Hedges d value of FC,COS and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Figure 7. Relationship of the Hedges   d of FC,COS with the variation in environmental conditions in the diurnal cycling groups. Subfingures (af) denote the linear regression between the Hedges d value of FC,COS and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Sustainability 14 02840 g007
Figure 8. Relationship of the Hedges   d of FC,CO2 with the variation in environmental conditions in the diurnal cycling groups. Subfigures (af) denote the linear regression between the Hedges d value of FC,CO2 and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Figure 8. Relationship of the Hedges   d of FC,CO2 with the variation in environmental conditions in the diurnal cycling groups. Subfigures (af) denote the linear regression between the Hedges d value of FC,CO2 and (a) Δlight intensity, the change in light intensity (μmol m−2 s−1); (b) Δtemperature, the change in environmental temperature (°C); (c) ΔrH, the change in rH (%); (d) ΔVPD, the change in vapor deficit (Pa); (e) ΔCa,CO2, the change in Ca,CO2 (ppt); (f) ΔCa,COS, the change in Ca,COS (ppm).
Sustainability 14 02840 g008
Figure 9. Relationships of the Hedges   d values for LRU to the Hedges   d values for FC,COS (a) and the FC,CO2 (b) in each diurnal cycling group.
Figure 9. Relationships of the Hedges   d values for LRU to the Hedges   d values for FC,COS (a) and the FC,CO2 (b) in each diurnal cycling group.
Sustainability 14 02840 g009
Figure 10. Relationships of the Hedges   d values for gs to the Hedges   d values for FC,COS (a) and FC,CO2 (b) in each diurnal cycling group.
Figure 10. Relationships of the Hedges   d values for gs to the Hedges   d values for FC,COS (a) and FC,CO2 (b) in each diurnal cycling group.
Sustainability 14 02840 g010
Figure 11. The relationships between the Hedges d of FC,COS and the Hedges d of FC,CO2 (a), as well as the Hedges d of LRU and the Hedges d of gs (b) in the diurnal cycling groups.
Figure 11. The relationships between the Hedges d of FC,COS and the Hedges d of FC,CO2 (a), as well as the Hedges d of LRU and the Hedges d of gs (b) in the diurnal cycling groups.
Sustainability 14 02840 g011
Table 1. Control (Contr.) and treatment (Treatm.) groups at each timescale.
Table 1. Control (Contr.) and treatment (Treatm.) groups at each timescale.
GroupTreatm.Contr.Treatm.Contr.Treatm.Contr.
Diel and SeasonalDailySeasonal
PAR ≤ 600 μmol m−2 s−1PAR > 600 μmol m−2 s−1Before DOY 171,
After DOY 266
DOY 171~266
DiurnalN vs. MEFA vs. DDD vs. N
6:00~9:00,
15:00~18:00
9:00~15:005:00~8:00,
17:00~20:00
8:00~11:00,
15:00~17:00
5:00~20:0020:00~5:00
ControlPAR ≤ 600 μmol m−2 s−1PAR > 600 μmol m−2 s−1
TotalAll groups
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, P.; Wang, S.; Chen, B.; Amir, M.; Wang, L.; Chen, J.; Ma, L.; Wang, X.; Liu, Y.; Zhu, K. Light and Water Conditions Co-Regulated Stomata and Leaf Relative Uptake Rate (LRU) during Photosynthesis and COS Assimilation: A Meta-Analysis. Sustainability 2022, 14, 2840. https://doi.org/10.3390/su14052840

AMA Style

Wang P, Wang S, Chen B, Amir M, Wang L, Chen J, Ma L, Wang X, Liu Y, Zhu K. Light and Water Conditions Co-Regulated Stomata and Leaf Relative Uptake Rate (LRU) during Photosynthesis and COS Assimilation: A Meta-Analysis. Sustainability. 2022; 14(5):2840. https://doi.org/10.3390/su14052840

Chicago/Turabian Style

Wang, Pengyuan, Shaoqiang Wang, Bin Chen, Muhammad Amir, Lei Wang, Jinghua Chen, Li Ma, Xiaobo Wang, Yuanyuan Liu, and Kai Zhu. 2022. "Light and Water Conditions Co-Regulated Stomata and Leaf Relative Uptake Rate (LRU) during Photosynthesis and COS Assimilation: A Meta-Analysis" Sustainability 14, no. 5: 2840. https://doi.org/10.3390/su14052840

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