Next Article in Journal
Tropospheric Ozone: A Critical Review of the Literature on Emissions, Exposure, and Health Effects
Next Article in Special Issue
Carbon Reduction Pathways for Chinese Provinces: Considering Carbon Intensity Convergence, Regional Development Plans, and Shared Socioeconomic Pathways
Previous Article in Journal
Trend Analysis and Spatial Source Attribution of Surface Ozone in Chaozhou, China
Previous Article in Special Issue
Spatiotemporal Dynamics of CO2 Emissions in China Based on Multivariate Spatial Statistics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbonyl Sulfide (COS) in Terrestrial Ecosystem: What We Know and What We Do Not

by
Jiaxin Li
1,2,
Lidu Shen
1,
Yuan Zhang
1,
Yage Liu
1,
Jiabing Wu
1 and
Anzhi Wang
1,*
1
CAS Key Laboratory of Forest Ecology and Silviculture, Institute of Applied Ecology, Chinese Academy of Sciences, Shenyang 110016, China
2
University of Chinese Academy of Sciences, Beijing 101408, China
*
Author to whom correspondence should be addressed.
Atmosphere 2024, 15(7), 778; https://doi.org/10.3390/atmos15070778
Submission received: 31 May 2024 / Revised: 17 June 2024 / Accepted: 21 June 2024 / Published: 28 June 2024
(This article belongs to the Special Issue Carbon Emission and Carbon Neutrality in China)

Abstract

:
Over the past six decades, carbonyl sulfide (COS) in terrestrial ecosystems has been extensively studied, with research focusing on exploring its ecological and environmental effects, estimating source–sink volume, and identifying influencing factors. The global terrestrial COS sink has been estimated to be about 1.194–1.721 Tg a−1, with the terrestrial sink induced by plants and soils 0.50–1.20 Tg a−1, accounting for 41%–69% of the total. Hence, the role of plants and soils as COS sinks has been extensively explored. Now we know that factors such as the activity of carbonic anhydrase (CA), leaf structural traits, soil microbial activity, and environmental factors play significant roles in the COS budget. Developments in observational techniques have also made important contributions to the COS budget. This paper provides an overview of the research progress made on COS based on a comprehensive review of the literature. Then, it highlights the current research hotspots and issues requiring further exploration. For instance, it has been demonstrated that there are still significant uncertainties in the estimation of COS sources and sinks, emphasizing the need for further exploration of COS measuring techniques. This review aims to provide comprehensive guidance for COS research in terrestrial ecosystems.

1. Introduction

Carbonyl sulfide (COS) is a colorless gas that resembles sulfur, with a boiling point of −50.2 °C. Its existence was first postulated by J.P. Couërbe [1] in the 17th century, and Carlvon provided more details of its characteristics thirty years later. In 1957, Ferm [2] conducted an extensive review of COS chemical properties, reactions, and preparation methods. During the 1960s, scientists initially identified COS as a minor trace gas mainly present in the troposphere of the global atmosphere. Due to its limited reactivity within the atmospheric layer, COS displays a relatively uniform distribution, with an average concentration of around 500 pptv, and can persist for one to four years [3].
Turco et al. [4] reported that the increasing emissions of COS from anthropogenic activities may have significant implications for climate change, and for the first time, the global source of COS was approximately 5 Tg a−1. Subsequently, substantial research was conducted concerning the dynamics of atmospheric COS associated with fossil fuel combustion, and numerous studies have been initiated to measure atmospheric concentrations of COS. Over the 20th century, research primarily focused on interannual budget imbalances and revealed substantial disparities between the sources and sinks of COS [5]. The total COS budget has been estimated in recent studies using the model inversion, yielding results that differ from previous research data. For example, in more recent research, Ma et al. [6] combined NOAA surface and MIPAS satellite observations to estimate the global budget of COS and reported that the biosphere uptake of COS was 922.9 ± 28.7 GgS a−1, a much smaller value compared to a prior value of 1053 GgS a−1 [7]. The continuous exchange of COS among the atmosphere, plants, and soil (Figure 1) suggests its potential role as a precursor in the terrestrial ecosystem cycling of carbon disulfide (CS2) [8]. In the early 2000s, Blake et al. [9] discovered that COS could effectively indicate carbon dioxide absorption during photosynthesis. Despite its concentration in the atmosphere being approximately six orders of magnitude lower than that of carbon dioxide (CO2) in the atmosphere, its “signal” is six times stronger than that of CO2, which helps to avoid complexities associated with the simultaneous measurement of CO2 uptake and release. Since then, COS has emerged as a reliable indicator of gross primary productivity (GPP) in vegetation.
Understanding and quantifying COS sources, sinks, and flux processes is crucial for global GPP estimation and the studies of global atmospheric chemical and biogeochemical cycles. Current research primarily focuses on the following: (1) elucidating the cooling effect of COS in the stratosphere and its role as a tracer in terrestrial ecology; (2) investigating mechanisms and influencing factors related to COS gas transport between the atmosphere, vegetation, and soil; (3) developing observation methods for COS fluxes and model predictions to improve accuracy in estimating COS sources and sinks. The main method used in this research is a systematic literature review. Based on a qualitative literature review, we hope to identify the current hotspots and the knowledge gaps in COS research and bring new insights for further investigation.

2. Ecological and Environmental Impacts of COS

2.1. Stratospheric Climate Impact

COS is a major source of stratospheric sulfate aerosols, which can affect atmospheric radiation balance and global climate change. It is one of the most abundant sulfur-containing gases in the atmosphere, with a global COS mole fraction ranging from 350 to 550 ppt (parts-per-trillion, i.e., 10−6 ppm). Due to its long atmospheric lifetime and low solubility, a significant portion of COS can reach the stratosphere at an altitude of 30 km [10], undergo photolysis under high levels of ultra-violet radiation, resulting in the formation of SO2, and eventually sulfate aerosol particles [11]. These aerosols serve as the primary source of stratospheric sulfate aerosols during non-volcanic eruption periods. They possess the potential to reflect short-wave radiation emitted by the sun and contribute to a negative direct solar radiative forcing, thereby exerting a significant impact on the ozone layer and other ecological factors. Notably, aerosols located in the lowermost stratosphere have a particularly significant impact. Studying aerosol characteristics and sources in this specific region under volcanic stasis conditions has revealed that fine sulfur production in the stratosphere amounts to 0.066 Tg a−1 [12,13]. Approximately half of this sulfur is produced by carbonyl sulfide located at the uppermost part of this lowest stratospheric layer. Stratospheric sulfate aerosols offer an interface for multiphase reactions within our atmosphere that contribute to ozone depletion, affect atmospheric radiation balance, and ultimately lead to global climate change. Therefore, investigating the heterogeneous reactions between carbonyl sulfur and atmospheric aerosols is crucial for revealing sulfate aerosol formation mechanisms and exploring the interconnection of COS with the global atmospheric environment.
However, the cooling effect of COS on sulfur aerosol formation in the stratosphere has initiated a new debate on mitigating global warming. At the end of the last century, some scientists proposed artificial COS emissions, which indirectly enhance stratospheric sulfur aerosol particles, as a means for counteracting global warming [14,15,16]. Visioni et al. [17] used GISS-E2.1-G and the CESM2-WACCM6 and UKESM1.0 models to demonstrate that it is possible to use these simulations to comprehensively inject COS for stratospheric strategies. Nevertheless, many other scientists believed that COS could warm up at the lower atmosphere surface and hence create a greenhouse effect that counteracts the cooling effect of sulfur-containing aerosols in the stratosphere [18,19]. According to the study conducted by Brühl et al. [11], COS exhibits absorption bands in the infrared (IR) spectrum and has a significant global warming potential, estimated to be 97 times that of CO2 over the next 20 years and 27 times over a 100-year scale. These findings indicate that COS significantly influences short-term climate change rather than long-term climate change.

2.2. Photosynthetic Tracer

COS has been confirmed as an effective tracer for GPP in various environments [20]. It is a sulfur-containing gas, like CO2, which can also be absorbed by plants. However, no one realized there was a direct connection between them until Goldan et al. [21] established a connection between the absorption of COS and CO2 by plants, suggesting that plant photosynthesis plays a crucial role in the uptake of both gases. They found that COS and CO2 diffuse from the atmosphere to the plant through stomates, but COS is hydrolyzed within the leaf under the catalytic effect of CA. Their absorption rates are positively correlated [22], whereas leaves do not produce COS; they emit CO2 through respiration [23,24]. This can explain why vegetation COS fluxes could be used as a proxy for GPP. The process of COS gas exchange in plants involves simultaneous production and consumption [25,26]. Plant respiration generally does not generate COS. Therefore, COS is identified as a promising proxy of GPP. Zhu et al. [27] simulate the differences in COS uptake between sunshade and sunlight leaves in the Boreal Ecosystem Productivity Simulator (BEPS), significantly improving the accuracy of GPP simulations in different ecosystems. Long-term changes in atmospheric COS have been influenced by significant increases in plant photosynthesis during the 20th century. Meanwhile, the role of COS as a photosynthesis tracer has also been employed to investigate the impact of extreme environments on plant photosynthetic processes. For instance, Sun et al. [28] used COS as a photosynthetic tracer, revealing that under heatwaves, the photosynthesis COS uptake is suppressed, leading to a lower internal leaf conductance to COS. Moreover, regional and leaf-scale variations in atmospheric COS concentration changes closely correspond to those of CO2 concentration changes. Comparing the global distribution and seasonal changes of COS with those of CO2 demonstrated that COS can serve as an indicative tracer for global CO2 flux [29]. Campbell et al. [30] discovered that the COS records derived from observations exhibited good consistency with climate and simulated carbon budget, thereby enabling the reconstruction of terrestrial vegetation productivity during the last ice age. This provides a solid foundation for investigating long-term variations in terrestrial ecosystem productivity. The measurement of COS concentration in ice cores offers a novel approach for studying the response of terrestrial ecosystems to climate change by splitting carbon sources and sinks in ecosystems through COS characterization, thus facilitating the examination of individual flux processes in reaction to climate change [31,32].

3. Source and Sink of COS

Since the 17th century, the concentration of carbonyl sulfur in the troposphere has increased by 28.7%, rising from 373 ppt in 1616–1694 to 480 ± 60 ppt at present [33]. However, this temporal evolution fluctuated significantly; for instance, tropospheric COS displayed a declining trend from 1995 to 2003 but gradually increased and stabilized from 2003 to 2015 before resuming a downward trend from 2015 to 2020 [34,35]. This phenomenon may be closely associated with the upsurge of sulfides, such as CS2 and COS, resulting from anthropogenic industrial emissions since the onset of the twenty-first century. The latest SiB4.2 models estimate global terrestrial ecosystem COS fluxes ranging between 0.73 and 1.5 Tg a−1 [7]; nevertheless, significant imbalances exist within global COS budgets concerning total source and sink budgets, as shown in Figure 2 [36,37,38,39,40]. Despite different study methods that may contribute to part of the discrepancy, there must still be undiscovered COS sources or sinks in the Earth system that we have not yet discovered. The following subsections explore the investigations and findings about the COS flux of plants and soils in various ecosystems, such as forests, grasslands, croplands, and wetlands.

3.1. Plant

The primary mechanism for removing COS from the atmosphere is through plant absorption, which is particularly pronounced during the growing season of terrestrial ecosystems [42], and the role of vegetation as a sink in the troposphere has been extensively studied for over 40 years. The estimated global terrestrial vegetation COS sink is 0.2–1.36 Tg a−1 [36,42,43,44,45].
Forests serve as significant sinks for COS, with the rainforest regions of South America, Africa, and Southeast Asia being the primary areas for COS uptake by plants globally. The absorption fluxes of COS by plant leaves play a crucial role in forest ecosystems and can be assessed by measuring profiles of COS within and above the canopy, reflecting the dynamics of COS gas exchange. Different tree species display varying levels of absorption fluxes for COS. While most studies suggest that trees act as atmospheric sinks for COS, some studies also reported conflicting findings. For instance, spruce forests exhibit sink effects during the day but predominantly release effects were observed at night, which led to these forests presenting no clear absorption on a daily basis [46]. Moreover, net release effects of COS have been observed in loblolly pine, Sophora sinensis, Salix amabilis, and white elm [47], highlighting the need for further research on various tree species to establish comprehensive data sets on their roles as sources or sinks of COS. Additionally, studying the physiological mechanisms underlying these specific responses within different plant species is essential.
The COS source and sink effect of grassland is primarily influenced by temperature and precipitation in different seasons, exhibiting greater fluctuations compared to forests. In spring, grasslands typically act as a source of COS emissions, with significantly higher average COS fluxes than in summer and autumn [48]. Whelan et al. [49] employed the static box method to observe COS fluxes in California’s grassland ecosystem during both growing and non-growing seasons, revealing that temperature is the main factor influencing COS exchange during the dry, non-growing season. However, during the rainy season, the sink capacity for COS within the system increased [50]. This dynamic feature can be used to characterize the effect of drought on grassland ecosystem productivity.
The COS source and sink capacity of different crops also varies with phonology. For instance, research has found that COS flux transitions from nearly zero to release as wheat grain heads transition from green to brown [43]. This may be attributed to the stimulation of COS production by sulfur-containing amino acids during grain filling and leaf aging, considering that the distribution of S protein in wheat changes at different growth stages [51]. Brown et al. [52] employed isotope labeling techniques and discovered that rice COS exchange rate increases with plant growth, peaking during the heading stage [53]. However, a significant portion of absorbed COS by crops is emitted into the atmosphere through straw incineration. Nguye et al. [54] analyzed atmospheric samples during straw burning in Vietnam and found an annual release of approximately 0.02 Tg of COS into the atmosphere, suggesting that rice straw burning could become a significant source of COS at the global scale.
Although vegetation is the primary global reservoir for COS, laboratory studies have revealed that lichens, green algae, and fungi also possess comparable COS absorption potential to vascular plants [55], exhibiting absorption rates ranging from 0.02 to 0.14 pmol g−1s−1. Field investigations have confirmed these findings, while the absorption rates observed in the field are slightly lower. The COS sink from these organisms amounts to approximately 0.3 Tg a−1 [56], which may not be considered a major sink but should not be ignored [57]. In regions where lichens dominate, atmospheric COS concentrations have been observed to decline. Air mass measurements conducted over cold regions in central and northern Canada exhibited a 19% reduction in atmospheric COS concentration within the air mass [58]. Lichen-dominated ecosystems encompass diverse land areas, particularly in northern and Arctic regions, as well as the Siberian tundra [59], potentially accounting for unidentified regions with notable COS sinks. However, considering that mossy lichen constitutes only 50% of vegetation cover in these regions alone cannot explain such substantial consumption of atmospheric COS [60]. Henceforth, there probably exist other potential sinks within the above-mentioned regions.

3.2. Soil

Soil uptake of COS represents the second largest sink for atmospheric COS, following plant uptake. Compared to the discovery of COS absorption by plants, scientists discovered very late that soil also exhibits source or sink effects. In the 1990s, it was believed that soil was one of the primary sources of atmospheric COS emissions, accounting for 21–25% of the total source [37]. However, subsequent field investigations revealed that soil also acts as a sink for COS [61], though its absorption capacity is significantly lower than that of plant leaves. Some researchers estimated that soil accounted for 26–55% of the global COS sink based on flux measurements [62]. Recent global estimation studies also suggest a potential shift in overall directionality from source to sink in terms of COS flux between soil and atmosphere [52,53].
Most forest soils act as sinks for COS [63]. However, during the spring season, COS emissions have been observed from broadleaf and pine tree forest soils in Harvard Forest [64]. COS absorption by the canopy is a major factor that modulates forest COS sink due to a small amount of COS absorption by forest soil. For instance, the range of COS flux in temperate and subtropical forest soils varies from −8 ± 1.45 pmol m−2 s−1, accounting for approximately 8% of the observed ecosystem-scale COS flux [65]. It is worth noting that there is currently a lack of data on COS flux between tropical and subtropical forest soils and the atmosphere, which hampers our understanding of global COS budgets within these systems. Additionally, interactions between nitrogen and sulfur are also a matter of concern. The research findings suggest that nitrogen addition significantly enhances COS emissions in broadleaf forests, acidic forest soils, and dryland soils. However, no significant impact of nitrogen addition on COS emissions is observed in pine forests; instead, it can stimulate the absorption of COS in northern coniferous forest soils [58].
The fluctuation amplitudes of grassland COS exchanges are more pronounced compared to forests. COS flux of grassland soil is influenced by radiation during the early growing season. Intense radiation on the exposed ground surface impacts the exchange of COS between the soil and atmosphere [66]. Agricultural soils, except for paddy fields, act as natural sinks for COS under normal conditions. However, net COS emissions can occur in all cropland soils with high water content during summer. Research has indicated that an anaerobic environment promotes COS production, and certain pathways for COS production are hindered under aerobic conditions [67], which may be attributed to differences in microbial community structure among different soils. When anaerobic microorganisms dominate the community, their optimal reaction conditions will inevitably be disrupted by aerobic environments.
Wetlands emit COS under hypoxic conditions via emergent plant stems. In contrast, non-wetland soils release it within anoxic regions of the soil profile or under aerobic conditions [68]. Changes in hydrological conditions, such as flooding duration and frequency, can impact the sulfur cycle in wetlands. Additionally, salt content is one of the factors that influences the flux of COS in wetland ecosystems. Plants produce COS as they regulate osmotic pressure in saline environments. Further investigation is needed to explore the relationship between net flux and temperature as well as soil moisture through field conditions [60].

4. Mechanism of COS Production/Consumption and Main Influencing Factors

The COS and CO2 molecules are transported to plant leaves and soil through the same physical diffusion pathway, where they undergo hydration reactions catalyzed by carbonic anhydrase (CA) to form H2S and CO2. Quantitative analysis revealed that leaf components associated with photosynthesis are the primary factor [69]. The leaves serve as major sinks for COS on a global scale, accounting for 62% of the total ecosystem sinks. The exchange of COS between the atmosphere and plants primarily occurs within the extracellular matrix and intercellular spaces located on the surface of plant leaves. The absorption of COS by plants can serve as an indicator of the physiological and ecological processes occurring within plants, considering that it enables the characterization of variations in stomatal conductance, mesophyll conductance, and CA activity among different plant species.

4.1. The Function of CA

CA has been identified as a pivotal enzyme involved in the consumption of COS in algae, lichens, higher plants, and soils [24,70]. During photosynthesis, COS is usually degraded by mediated CA because CA and ribulose-1,5-bisphosphate carboxylase (Rubisco) enzymes can effectively catalyze the decomposition of COS. While assimilating CO2, CA promotes COS consumption in leaves and direct conversion into CO2 and hydrogen sulfide (H2S) [41]. Reversible hydration reactions have been observed during this process. However, they predominantly proceed in one direction [71]. A portion of the produced H2S is further transformed into organic sulfur compounds for protein synthesis, including cysteine [72], while the remaining fraction is released from leaves into the atmosphere, serving as a significant source of atmospheric H2S [70]. By monitoring the consumption activity of CA on both CO2 and COS in the absence of other enzymes, it has been discovered that CA exhibits a higher affinity for COS than CO2 [73]. Moreover, at pH = 8.2, it can accelerate the consumption of COS [24], indicating a strong dependence on CA activity for COS absorption. Nevertheless, inhibiting CA only leads to a 50% reduction in COS acquisition [74], suggesting the potential involvement of non-inhibitory or non-enzymatic decomposition processes.
Different photosynthetic types also result in variations in the pathways of CA, which promotes COS consumption. Photosynthesis in higher plants includes C3, C4, and CAM metabolism. C4 plants contribute to approximately 18–25% of global photosynthesis [75]. The absorption of COS by C3 plants is attributed to the activity of CA, which facilitates the diffusion of CO2 into the chloroplast mechanism of all photosynthetic cells [76]. The inherent capacity of CA is strongly correlated with the rate of COS absorption, and COS undergoes direct metabolism by CA after traversing the chloroplast membrane in C3 plants. The CO2 absorption efficiency of C4 plants is higher, however, the CA activity in C4 plants generally exhibits lower levels compared to that of C3 plants [77]. Consequently, the resistance towards COS absorption may be greater in C4 plants than in C3 plants. The concentration of CA in the leaves of cultivated sorghum, a C4 plant, is relatively low and primarily localized within the cytoplasm of mesophyll cells characterized by large intercellular spaces [78]. Within the cytoplasm of mesophyll cells in C4 plants, CA enzymes directly hydrolyze COS to CO2 or HCO3-, which are subsequently transported to bundle sheath cells via malic acid and aspartic acid, respectively, providing energy for the Calvin cycle [73] (Figure 3).
Although the catalytic of CA on the decomposition of COS is a very important process during photosynthesis, its function is independent of light conditions. In vascular plants, photosynthetic CO2 assimilation is reduced under low light conditions, while the CA is independent of light, i.e., the activity of the CA remains unaffected during night [78], allowing the plant to continue absorbing COS. Wehr et al. [79] found that considering CA as constant over an extended time scale enables the utilization of COS as a valuable tool in studying stomatal conductance during down-scaling studies.
The primary role of CA in soil COS consumption is predominantly attributed to its presence within the microbial community. Isotopic studies have confirmed the existence of CA in various soils, and it plays an active role in the hydrolysis of COS by soil microorganisms. It has been demonstrated that CA exhibits activity in consuming COS within microorganisms [80]. Different types of CA exhibit varying affinities for COS, and the expression of CA activity information differs among different microbial communities. Measurement of COS flux helps to understand how enzyme activity and root metabolism influence the microbial decomposition of COS, thereby determining the complete dynamic process of soil COS flux, which can be further parameterized when combined with a model. The COS is more likely to reach the reactive enzyme site after cell membrane lysis, and factors such as extracellular metabolic characteristics of respiratory enzymes contribute to the persistence of COS consumption even after soil microbial inactivation treatment [81,82,83]. This suggests that apart from microbial action, enzyme stability plays a significant role in influencing the COS exchange process.

4.2. Leaf Structural Traits

Leaf structural traits mainly include leaf vein characteristics, stomatal features, the anatomical structure of leaf cross-sectional tissues, and other leaf anatomical traits [84]. The opening and closing of the stomatal is considered one of the important factors affecting the exchange of COS gas between plants and the environment [85]. The leaf stomata feature provides a basis for COS conductance studies, this leaf-to-canopy upscaling approach will improve the accuracy of GPP estimates on regional to global scales [28,86]. Stimler et al. [71] found a correlation between the exchange of COS and the overall absorption of CO2, with stomata playing a regulatory role in this process. During photosynthesis in daylight, COS enters mesophyll cells through stomata. The rate of COS uptake by plants is proportional to the stomatal conductance; however, during the night, some plants partially close their stomata, leading to the release of COS into the environment through the stomata. This phenomenon was discovered in both C3 and C4 plants as early as the end of the 19th century [87]. After that, the phenomenon of stomata remaining open throughout the night was also observed in Arabidopsis mutants [88]. The release of COS during the night may transform plants from daytime sinks to nighttime sources. While most experiments on stomatal conductance in plants are typically conducted during daylight, the regulation of nocturnal stomatal activity has significant ecological implications.
The effect of stomatal conductance on COS flux is also affected by temperature fluctuations in different seasons. High temperature in summer causes plants to release COS, and meanwhile, the leaf renewal rate in summer is higher than that in autumn, and the average COS absorption rate of young leaves is larger; while in winter with lower temperature, both stomatal opening and enzyme activity are weak, which reduces the absorption of COS by plants [47]. Therefore, studies on the mechanisms of plant COS fluxes, such as changes in stomatal openings at the leaf scale and the effects of changes in metabolic processes on COS fluxes due to senescence, can characterize seasonal fluctuations in COS fluxes in terrestrial ecosystems.

4.3. Soil Condition

The settlement rate of COS on the soil surface is influenced by various factors, including soil type, nutrient content, soil temperature, soil moisture, and atmospheric COS mixing ratio [89]. Both biological and abiotic processes contribute to the overall exchange of COS, exhibiting variations across different soils. These multiple factors cause considerable uncertainty when studying the sources and sinks of COS in soils. The production and consumption of COS in soil exhibit a strong dependence on temperature and humidity. For example, COS emissions from the soil occur when the temperature approaches 0 °C or 15–20 °C [90]. Enzymatic catalysis significantly influences COS emission at temperatures between 15 and 20 °C, thereby promoting its release. Moreover, fluctuations in temperature can induce the desorption of COS from soil minerals [91]. COS is released when soil moisture reaches extreme levels, with peak emissions occurring under very low soil humidity conditions. The absorption of COS takes place within a range of moderate humidity, and the release phenomenon of COS is also observed in anaerobic environments resulting from super-saturation [92]. Changes in soil moisture caused by water stress can trigger pulse emissions of COS, and lower water content may lead to a decrease in seasonal COS flux. This could be attributed to the impact of soil water content on pore filling, which affects the diffusion of COS within the soil [93]. Furthermore, this variation can influence both the diversity and function of soil biological communities as different microbial communities exhibit activity under distinct soil moisture conditions [94,95]. Simultaneously, microbial activity and environmental adaptability differ across various soil conditions. Alterations in microbial community composition can affect organic matter decomposition in soils, indirectly resulting in a stronger response of COS flux to UV radiation exposure in soils with higher organic matter content [96].

4.4. Microbial Action

The studies have revealed that COS can serve as a reliable indicator of plant disturbance. For example, COS emission has been observed in rape plants after fungal infection, which may be linked to the plant stress response [85]. The leaf or interleaf microbial community acts as the interface for plant–fungi interactions, and fungal infection induces alterations in COS absorption within plant leaves.
Over the past century, research on COS in soil has been primarily focused on the source–sink relationship at a macroscopic flux level, with limited exploration of the underlying mechanisms of COS transformation. Soil microorganisms play a crucial role in mediating trace gas exchange between the soil and atmosphere. The consumption of COS by soil is dependent on microbial activity within the soil environment. Environmental COS concentration, as well as soil physical and chemical properties, temperature, and other physical factors, indirectly influence microbial activity.
COS can serve as a nutrient source for both autotrophic and heterotrophic bacteria. Bacteria and fungi are the predominant microbial communities in soil, with fungi playing a significant role in COS absorption [97], particularly in forest soils, where they act as important converters. The initial discovery of fungal degradation of COS was achieved during the uptake of COS by Fusarium solani strain THIF01 under chemoautotrophic conditions [98]. Purified saprophytic fungi such as Fusarium solani and Trichoderma spp have been found to consume atmospheric COS, while Mucoromycotina fungi can release COS. Sordariomycetes fungi are also capable of degrading COS. The impact of bacteria on COS has been more extensively studied compared to that of fungi, with several soil microorganisms showing depletion effects on the CA-induced consumption of COS. Thiobacillus, Mycobacterium Dietzia maris NBRC15801T, Streptomyces ambofaciens NBRC12836T, Actinomycetales bacteria, and Streptomyces have all demonstrated the ability to both release and degrade COS after inoculation into sterilized soil [99,100]. Additionally, a highly specific enzyme responsible for degrading COS has been isolated from Thiobacillus thioparus strain THI115 [101]. Behrendt et al. [102] first evaluated the environmental significance of OCS production and consumption by different microbial groups through various enzymes other than CA to gain insight into the mechanisms underlying microbial consumption and production of COS in soil.

5. Conclusions and Perspective

To summarize, research on COS has experienced a surge in recent years, covering various aspects such as exchange mechanisms among different vegetation and soil types, multiple observation methods in diverse ecosystems, and retrieval of historical atmospheric records from ice cores dating back to the Last Glacial Maximum. The utilization of aircraft and satellites for monitoring COS dynamics at regional or global scales has enabled the development of climate models. Multi-scale observation integration has elevated the precision of COS source–sink evaluations in different regions and enhanced the understanding of the underlying mechanisms of its formation and turnover processes. However, more observations and investigations are still needed to characterize COS sources, sinks, and flux processes, as well as their ecological implications at different spatiotemporal scales, especially their links to stratospheric climate.

5.1. The Ecological and Environmental Impacts of COS Are Becoming Increasingly Evident

The exploration of atmospheric COS has progressed from examining its lifetime and chemical properties in atmospheres to investigating its transport pathway and climatic impacts. The total COS flux from the troposphere to the stratosphere is 0.137 Tg a−1, the stratospheric sink is about 0.032–0.078 Tg a−1, and the remaining 0.082 Tg a−1 is presumed to return from the stratosphere to the troposphere [10]. Around 60% of COS returned to the troposphere, and part of it may remain in the lower atmosphere as a greenhouse gas, resulting in the warming effect of COS in the stratosphere and the cooling effect of COS forming sulfur aerosols in the stratosphere tending to offset each other. Another part of COS is absorbed by vegetation in the lower atmosphere. However, further investigation is necessary to determine the exact amount of COS gases remaining in the troposphere as greenhouse gases, and how much of them enter the terrestrial ecosystem circulation. Additionally, there is a need to further explore the mechanism of how COS forms aerosols in the stratosphere.
By partitioning the COS fluxes originating from vegetation and soil, we can quantify the CO2 fluxes associated with photosynthesis and respiration of plants, as well as assess the impacts of changes in atmospheric CO2 concentrations on vegetation. The magnitude of COS signals observed through satellite and regional-scale atmospheric measurements is six times greater than that of CO2, indicating that COS presents a distinct advantage in terms of detectability compared to CO2. Consequently, measuring atmospheric levels of COS provides an alternative method to accurately quantify both sources and sinks of CO2. Specifically, taking COS as a tracer enables us to establish a novel approach for estimating global GPP by assessing its contribution to carbon uptake. Extensive studies conducted across various scales, from individual plant leaves to regional spatial extents, consistently demonstrated the potential of COS as an effective indicator for GPP. Discrepancies between the current global and regional GPP observations and the predicted results indicate significant uncertainties in GPP models [103,104]. Introducing the COS absorption of vegetation into models can partially improve model performance. Since there is a mixing between COS soil and vegetation fluxes, a further development will be to refine the estimation for COS soil fluxes to provide a clearer delineation of COS fluxes between vegetation and soil. Environmental stress poses additional challenges for predicting regional and global COS distribution and, hence, GPP modeling. Therefore, the acquisition of more field observation data is also essential to improve the accuracy of regional or global GPP estimation models.

5.2. The Estimation Methods of COS Require Further Improvement

The assessment of COS sources and sinks is crucial for estimating GPP in terrestrial ecosystems. Previous studies primarily focused on annual budget closure, assuming a distinct spatial division between COS sources and sinks. However, as shown in Figure 2, the estimated global COS budget presents considerable discrepancy. Advancements in measurement technology have revealed that additional sources and sinks exist in terrestrial ecosystems, posing new challenges to the utilization of COS as a tracer for direct GPP measurement. Furthermore, it should be noted that COS sources and sinks also exhibit temporal variations, with the seasonal variation of oceanic COS sources being less prominent compared to that of terrestrial plant sinks. This discrepancy contributes to the differences in atmospheric COS observations between the northern and southern hemispheres. Therefore, when estimating global COS sources and sinks, we must consider the influence of dynamic processes governing atmospheric COS flow on source-sink coupling. For atmospheric measurements close to urban areas, the influence of anthropogenic sources of COS on the measurement results should be considered.
There are numerous methodologies available for estimating the sources and sinks of COS in ecosystems at various scales. In recent years, spectral technology has emerged as a novel approach for monitoring atmospheric trace gases. For instance, the Fourier transform infrared (FTIR) spectroscopy method was employed to collect and invert regional atmospheric COS data, revealing a significant positive trend in atmospheric COS levels from 2001 to 2015 [34,35]. The development of quantum cascade lasers (QCLs) represents a pivotal technological advancement in high-precision mid-infrared spectrometry for the measurement of trace gases in the atmosphere. Recent research has demonstrated that CW QCLs exhibit narrower line widths and improved stability compared to traditional QCLs, thereby enhancing the performance of high-sensitivity and high-precision trace gas measurements [105]. In terms of monitoring ecosystem scales, ground-based FTIR spectroscopy inversion is particularly sensitive to low-altitude radiation and allows direct observation of changes within specific ecosystems. On a larger scale, satellite remote sensing offers enhanced sensitivity towards COS concentrations within the middle troposphere, upper troposphere, and stratosphere, providing extensive data on COS distribution [106,107]. Additionally, both the Global Monitoring Network and Atmospheric Sampling Program encompass synchronous data on COS and CO2 concentrations within vegetated areas’ free troposphere and atmospheric boundary layer [40].
Model simulation currently serves as a primary approach in investigating vegetation and soil COS fluxes and, thus, is also a hotspot in COS research. For instance, Goldan et al. [43] employed the GEOS-Chem CTM chemical transport model in conjunction with estimated COS fluxes to simulate seasonal variations in atmospheric COS levels and vegetation absorption, which improved the seasonal change modeling for both hemispheres. Serio et al. [108], using EMD analysis, successfully separated seasonal cycles from other influencing factors, quantified the trend of atmospheric COS variation, and identified a strong annual cycle for this compound. Remaud et al. [109] developed a climate inverse model to assess the potential and limitations of global total vegetation primary productivity, plant respiration, and COS emissions; however, they found that tropical regions were poorly constrained by this model due to limited observational data for validation. Similarly, utilizing an inversion technique with the TM5-3DVAR chemical transport model (TM5) combined with the SiB4 biosphere model [110], Zumkehr et al. [111] also discovered missing sources of COS in tropical regions through its absorption by soil and vegetation. Vesala et al. [112], focusing on evergreen coniferous forests (ENFs) in the northern hemisphere, verified that the absence of a sink for COS at high latitudes is closely linked to its absorption by evergreen coniferous forests. Some studies have constructed mechanism models for accessing the source and sink of COS intensity and seasonal dynamic changes [39,105,107], while all these models excluded the influence of water stress on soil gas fluxes [40]. The current modeling system requires more extensive research on the mechanisms underlying soil absorption of COS, which necessitates an update in measuring instruments to meet the accuracy m0easurement of COS exchanges.
Although the methods summarized above improve the accuracy of estimating COS sources and sinks, certain issues remain unresolved. The spatial scale of existing observational areas is limited, with a lack of observations in tropical ecosystems that contribute roughly 60% to global GPP. Establishing a large-scale COS flux observation tower within tropical forest regions would effectively deal with the challenge of observing climate-scale changes, provide a sufficient database for global COS data modeling, and subsequently supply data for predicting global COS sources that are currently missing. Inverting seasonal dynamic characteristics of the mixed ratio of COS from station measurements and atmospheric total column measurements within the COS database is a promising method. Moreover, there has been insufficient research conducted on the impact of different biological types on COS, necessitating more year-round measurements for currently underrepresented communities to lead to better top-down estimations of terrestrial ecosystem COS fluxes. It is also crucial to consider regional limitations in models as well as the coupling effect among multiple factors in future studies.

5.3. The Metabolic Mechanism and Influencing Factors of COS Require Further Investigation

The physiological and biochemical processes of COS in various environments establish the scientific foundation for elucidating the source–sink dynamics within ecosystems. CA, as a crucial enzyme that facilitates COS consumption by higher plants, mechanistically explains the sink function of COS [113]. Additionally, CA elucidates, to some extent, how increasing atmospheric CO2 concentrations influence the vegetation’s ability to absorb COS.
Early studies on COS absorption by plants primarily focused on elucidating the mechanisms occurring within plant leaves. Taylor et al. [114] discovered that the intercellular diffusion rate of COS within leaves was predominantly influenced by the gas’s water solubility in cell fluid and chemical reaction conditions. However, this explanation does not fully account for all variations observed in COS flux within the leaf blade. The study conducted by Krebs et al. [81] revealed the significant role of CA in promoting COS hydrolysis at pH = 8.2 and effectively distinguished the process differences of CA in promoting COS hydrolysis between C3 and C4 plants. The internal resistance to COS absorption varies among different plant species, which is a crucial factor influencing the overall COS flux. Additionally, the relative leaf absorption (LRU) of COS differs across plant species [115], primarily influenced by factors such as leaf area index, stomatal conductance, and leaf age. For instance, changes in leaf area index (LAI) play a vital role in affecting COS flux within grassland ecosystems. Understanding the action mechanism of stomata on COS and determining the specific activity of CA towards COS (typically measured as activity units per milligram of enzyme) [32] are currently prominent areas of research in plants concerning carbonyl sulfide exchange. Additionally, further field experiments are required to validate the methodology for studying canopy and stomatal conductance using carbonyl sulfide.

Author Contributions

Conceptualization, J.L. and Y.Z.; formal analysis, L.S.; investigation, Y.L.; data curation, J.L.; writing-original draft preparation, J.L., Y.Z. and Y.L.; writing-review and editing, J.W.; visualization, J.L. and L.S.; supervision, A.W.; project administration, J.W.; funding acquisition, J.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant Number: 32271873, 32171873).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Couërbe, J.P. Ueber Den Schwefelkohlenstoff. J. Prakt. Chem. 1841, 23, 83–124. [Google Scholar] [CrossRef]
  2. Ferm, R.J. The Chemistry Of Carbonyl Sulfide. Chem. Rev. 1957, 57, 621–640. [Google Scholar] [CrossRef]
  3. Torres, A.L.; Maroulis, P.J.; Goldberg, A.B.; Bandy, A.R. Atmospheric OCS Measurements on Project Gametag. J. Geophys. Res. 1980, 85, 7357–7360. [Google Scholar] [CrossRef]
  4. Turco, R.P.; Whitten, R.C.; Toon, O.B.; Pollack, J.B.; Hamill, P. OCS, Stratospheric Aerosols and Climate. Nature 1980, 283, 283–285. [Google Scholar] [CrossRef]
  5. Lennartz, S.T.; Marandino, C.A.; Von Hobe, M.; Cortes, P.; Quack, B.; Simo, R.; Booge, D.; Pozzer, A.; Steinhoff, T.; Arevalo-Martinez, D.L.; et al. Direct Oceanic Emissions Unlikely to Account for the Missing Source of Atmospheric Carbonyl Sulfide. Atmos. Chem. Phys. 2017, 17, 385–402. [Google Scholar] [CrossRef]
  6. Ma, J.; Kooijmans, L.M.J.; Glatthor, N.; Montzka, S.A.; Von Hobe, M.; Röckmann, T.; Krol, M.C. Combined Assimilation of NOAA Surface and MIPAS Satellite Observations to Constrain the Global Budget of Carbonyl Sulfide. Atmos. Chem. Phys. 2024, 24, 6047–6070. [Google Scholar] [CrossRef]
  7. Kooijmans, L.M.J.; Cho, A.; Ma, J.; Kaushik, A.; Haynes, K.D.; Baker, I.; Luijkx, I.T.; Groenink, M.; Peters, W.; Miller, J.B.; et al. Evaluation of Carbonyl Sulfide Biosphere Exchange in the Simple Biosphere Model (SiB4). Biogeosciences 2021, 18, 6547–6565. [Google Scholar] [CrossRef]
  8. Ren, Y. Is Carbonyl Sulfide a Precursor for Carbon Disulfide in Vegetation and Soil? Interconversion of Carbonyl Sulfide and Carbon Disulfide in Fresh Grain Tissues in Vitro. J. Agric. Food Chem. 1999, 47, 2141–2144. [Google Scholar] [CrossRef] [PubMed]
  9. Blake, N.J.; Campbell, J.E.; Vay, S.A.; Fuelberg, H.E.; Huey, L.G.; Sachse, G.; Meinardi, S.; Beyersdorf, A.; Baker, A.; Barletta, B.; et al. Carbonyl Sulfide (OCS): Large-scale Distributions over North America during INTEX-NA and Relationship to CO2. J. Geophys. Res. Atmos. 2008, 113, 2007JD009163. [Google Scholar] [CrossRef]
  10. Karu, E.; Li, M.; Ernle, L.; Brenninkmeijer, C.A.M.; Lelieveld, J.; Williams, J. Carbonyl Sulfide (OCS) in the Upper Troposphere/Lowermost Stratosphere (UT/LMS) Region: Estimates of Lifetimes and Fluxes. Geophys. Res. Lett. 2023, 50, e2023GL105826. [Google Scholar] [CrossRef]
  11. Brühl, C.; Lelieveld, J.; Crutzen, P.J.; Tost, H. The Role of Carbonyl Sulphide as a Source of Stratospheric Sulphate Aerosol and Its Impact on Climate. Atmos. Chem. Phys. 2012, 12, 1239–1253. [Google Scholar] [CrossRef]
  12. Martinsson, B.G.; Nguyen, H.N.; Brenninkmeijer, C.A.M.; Zahn, A.; Heintzenberg, J.; Hermann, M.; Van Velthoven, P.F.J. Characteristics and Origin of Lowermost Stratospheric Aerosol at Northern Midlatitudes under Volcanically Quiescent Conditions Based on CARIBIC Observations. J. Geophys. Res. Atmos. 2005, 110, 2004JD005644. [Google Scholar] [CrossRef]
  13. Crutzen, P.J. The Possible Importance of CSO for the Sulfate Layer of the Stratosphere. Geophys. Res. Lett. 1976, 3, 73–76. [Google Scholar] [CrossRef]
  14. Quaglia, I.; Visioni, D.; Pitari, G.; Kravitz, B. An Approach to Sulfate Geoengineering with Surface Emissions of Carbonyl Sulfide. Atmos. Chem. Phys. 2022, 22, 5757–5773. [Google Scholar] [CrossRef]
  15. Taubman, S.J.; Kasting, J.F. Carbonyl Sulfide: No Remedy for Global Warming. Geophys. Res. Lett. 1995, 22, 803–805. [Google Scholar] [CrossRef]
  16. Crutzen, P.J. Albedo Enhancement by Stratospheric Sulfur Injections: A Contribution to Resolve a Policy Dilemma? Clim. Chang. 2006, 77, 211. [Google Scholar] [CrossRef]
  17. Visioni, D.; Bednarz, E.M.; Lee, W.R.; Kravitz, B.; Jones, A.; Haywood, J.M.; MacMartin, D.G. Climate Response to Off-Equatorial Stratospheric Sulfur Injections in Three Earth System Models-Part 1: Experimental Protocols and Surface Changes. Atmos. Chem. Phys. 2023, 23, 663–685. [Google Scholar] [CrossRef]
  18. Von Hobe, M.; Brühl, C.; Lennartz, S.T.; Whelan, M.E.; Kaushik, A. Comment on “An Approach to Sulfate Geoengineering with Surface Emissions of Carbonyl Sulfide” by Quaglia et al. (2022). Atmos. Chem. Phys. 2023, 23, 6591–6598. [Google Scholar] [CrossRef]
  19. Pitari, G.; Aquila, V.; Kravitz, B.; Robock, A.; Watanabe, S.; Cionni, I.; De Luca, N.; Di Genova, G.; Mancini, E.; Tilmes, S. Stratospheric ozone response to sulfate geoengineering: Results from the Geoengineering Model Intercomparison Project (GeoMIP). Geophys. Res. Atmos. 2014, 119, 2629–2653. [Google Scholar] [CrossRef]
  20. Asaf, D.; Rotenberg, E.; Tatarinov, F.; Dicken, U.; Montzka, S.A.; Yakir, D. Ecosystem Photosynthesis Inferred from Measurements of Carbonyl Sulphide Flux. Nat. Geosci. 2013, 6, 186–190. [Google Scholar] [CrossRef]
  21. Goldan, P.D.; Kuster, W.C.; Albritton, D.L.; Fehsenfeld, F.C. The Measurement of Natural Sulfur Emissions from Soils and Vegetation: Three Sites in the Eastern United States Revisited. J. Atmos. Chem. 1987, 5, 439–467. [Google Scholar] [CrossRef]
  22. 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]
  23. 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 Hydrosulfide Analogue. Chembiochem 2007, 8, 530–536. [Google Scholar] [CrossRef] [PubMed]
  24. 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]
  25. Conrad, R. Soil Microbial Processes and the Cycling of Atmospheric Trace Gases. Philos. Trans. R. Soc. Lond. Ser. A Phys. Eng. Sci. 1995, 351, 219–230. [Google Scholar] [CrossRef]
  26. Remde, A.; Slemr, F.; Conrad, R. Microbial Production and Uptake of Nitric Oxide in Soil. FEMS Microbiol. Lett. 1989, 62, 221–230. [Google Scholar] [CrossRef]
  27. Zhu, H.; Xing, X.; Wu, M.; Ju, W.; Jiang, F. Optimizing the Terrestrial Ecosystem Gross Primary Productivity Using Carbonyl Sulfide (COS) within a “Two-Leaf” Modeling Framework. EGUsphere 2024, 2024, 1–34. [Google Scholar] [CrossRef]
  28. Sun, W.; Maseyk, K.; Lett, C.; Seibt, U. Restricted Internal Diffusion Weakens Transpiration–Photosynthesis Coupling during Heatwaves: Evidence from Leaf Carbonyl Sulphide Exchange. Plant Cell Environ. 2024, 47, 1813–1833. [Google Scholar] [CrossRef] [PubMed]
  29. 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 Similarities to CO2. J. Geophys. Res. 2007, 112, D09302. [Google Scholar] [CrossRef]
  30. Campbell, J.E.; Berry, J.A.; Seibt, U.; Smith, S.J.; Montzka, S.A.; Launois, T.; Belviso, S.; Bopp, L.; Laine, M. Large Historical Growth in Global Terrestrial Gross Primary Production. Nature 2017, 544, 84–87. [Google Scholar] [CrossRef]
  31. 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]
  32. Brugnoli, E.; Calfapietra, C. Carbonyl Sulfide: A New Tool for Understanding the Response of the Land Biosphere to Climate Change. New Phytol. 2010, 186, 783–785. [Google Scholar] [CrossRef] [PubMed]
  33. 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, e2020060118. [Google Scholar] [CrossRef] [PubMed]
  34. Kremser, S.; Jones, N.B.; Palm, M.; Lejeune, B.; Wang, Y.; Smale, D.; Deutscher, N.M. Positive Trends in Southern Hemisphere Carbonyl Sulfide: TRENDS IN SOUTHERN HEMISPHERE OCS. Geophys. Res. Lett. 2015, 42, 9473–9480. [Google Scholar] [CrossRef]
  35. Lejeune, B.; Mahieu, E.; Vollmer, M.K.; Reimann, S.; Bernath, P.F.; Boone, C.D.; Walker, K.A.; Servais, C. Optimized Approach to Retrieve Information on Atmospheric Carbonyl Sulfide (OCS) above the Jungfraujoch Station and Change in Its Abundance since 1995. J. Quant. Spectrosc. Radiat. Transf. 2017, 186, 81–95. [Google Scholar] [CrossRef]
  36. Ma, J.; Kooijmans, L.M.J.; Cho, A.; Montzka, S.A.; Glatthor, N.; Worden, J.R.; Kuai, L.; Atlas, E.L.; Krol, M.C. Inverse Modelling of Carbonyl Sulfide: Implementation, Evaluation and Implications for the Global Budget. Atmos. Chem. Phys. 2021, 21, 3507–3529. [Google Scholar] [CrossRef]
  37. Chin, M.; Davis, D.D. Global Sources and Sinks of OCS and CS 2 and Their Distributions. Glob. Biogeochem. Cycles 1993, 7, 321–337. [Google Scholar] [CrossRef]
  38. Suntharalingam, P.; Kettle, A.J.; Montzka, S.M.; Jacob, D.J. Global 3-D Model Analysis of the Seasonal Cycle of Atmospheric Carbonyl Sulfide: Implications for Terrestrial Vegetation Uptake. Geophys. Res. Lett. 2008, 35. [Google Scholar] [CrossRef]
  39. Berry, J.; Wolf, A.; Campbell, J.E.; Baker, I.; Blake, N.; Blake, D.; Denning, A.S.; Kawa, S.R.; Montzka, S.A.; 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]
  40. Launois, T.; Peylin, P.; Belviso, S.; Poulter, B. A New Model of the Global Biogeochemical Cycle of Carbonyl Sulfide-Part 2: Use of Carbonyl Sulfide to Constrain Gross Primary Productivity in Current Vegetation Models. Atmos. Chem. Phys. 2015, 15, 9285–9312. [Google Scholar] [CrossRef]
  41. 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]
  42. Maseyk, K.; 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] [CrossRef] [PubMed]
  43. Goldan, P.D.; Fall, R.; Kuster, W.C.; Fehsenfeld, F.C. Uptake of COS by Growing Vegetation: A Major Tropospheric Sink. J. Geophys. Res. 1988, 93, 14186–14192. [Google Scholar] [CrossRef]
  44. 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, 4759. [Google Scholar] [CrossRef]
  45. Hu, L.; Montzka, S.A.; Kaushik, A.; Andrews, A.E.; Sweeney, C.; Miller, J.; Baker, I.T.; Denning, S.; Campbell, E.; Shiga, Y.P.; et al. 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] [PubMed]
  46. 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. 2002, 2, 171–181. [Google Scholar] [CrossRef]
  47. Geng, C.; Mu, Y. Carbonyl Sulfide and Dimethyl Sulfide Exchange between Trees and the Atmosphere. Atmos. Environ. 2006, 40, 1373–1383. [Google Scholar] [CrossRef]
  48. Geng, C. Carbonyl Sulfide and Dimethyl Sulfide Exchange between Lawn and the Atmosphere. J. Geophys. Res. 2004, 109, D12302. [Google Scholar] [CrossRef]
  49. Whelan, M.E.; Rhew, R.C. Reduced Sulfur Trace Gas Exchange between a Seasonally Dry Grassland and the Atmosphere. Biogeochemistry 2016, 128, 267–280. [Google Scholar] [CrossRef]
  50. Hoover, D.L.; Rogers, B.M. Not All Droughts Are Created Equal: The Impacts of Interannual Drought Pattern and Magnitude on Grassland Carbon Cycling. Glob. Chang. Biol. 2016, 22, 1809–1820. [Google Scholar] [CrossRef]
  51. Zhao, F.; Hawkesford, M.; McGrath, S. Sulphur Assimilation and Effects on Yield and Quality of Wheat. J. Cereal Sci. 1999, 30, 1–17. [Google Scholar] [CrossRef]
  52. Brown, K.A.; Kluczewski, S.M.; Bell, J.N.B. Metabolism of [35S]-Carbonyl Sulphide in Perennial Ryegrass (Lolium perenne L.) and Radish (Raphanus sativus L.). Environ. Exp. Bot. 1986, 26, 355–364. [Google Scholar] [CrossRef]
  53. Yi, Z.; Wang, X.; Sheng, G.; Fu, J. Exchange of Carbonyl Sulfide (OCS) and Dimethyl Sulfide (DMS) between Rice Paddy Fields and the Atmosphere in Subtropical China. Agric. Ecosyst. Environ. 2008, 123, 116–124. [Google Scholar] [CrossRef]
  54. Nguyen, B.C.; Mihalopoulos, N.; Putaud, J.P. Rice Straw Burning in Southeast Asia as a Source of CO and COS to the Atmosphere. J. Geophys. Res. 1994, 99, 16435–16439. [Google Scholar] [CrossRef]
  55. Gries, C.; Nash, T.H.; Kesselmeier, J. Exchange of Reduced Sulfur Gases between Lichens and the Atmosphere. Biogeochemistry 1994, 26, 25–39. [Google Scholar] [CrossRef]
  56. Kuhn, U.; Kesselmeier, J. Environmental Variables Controlling the Uptake of Carbonyl Sulfide by Lichens. J. Geophys. Res. Atmos. 2000, 105, 26783–26792. [Google Scholar] [CrossRef]
  57. Kuhn, U.; Wolf, A.; Gries, C.; Iii, T.H.N.; Kesselmeier, J. Field Measurements on the Exchange of Carbonyl Sulfide between Lichens and the Atmosphere. Atmos. Environ. 2000, 34, 4867–4878. [Google Scholar] [CrossRef]
  58. Johnson, J.E.; Bandy, A.R.; Thornton, D.C.; Bates, T.S. Measurements of Atmospheric Carbonyl Sulfide during the NASA Chemical Instrumentation Test and Evaluation Project: Implications for the Global COS Budget. J. Geophys. Res. 1993, 98, 23443. [Google Scholar] [CrossRef]
  59. Kershaw, K.A. The Role of Lichens in Boreal Tundra Transition Areas. Bryologist 1978, 81, 294–306. [Google Scholar] [CrossRef]
  60. Fried, A.; Klinger, L.F.; Erickson, D.J. Atmospheric Carbonyl Sulfide Exchange in Bog Microcosms. Geophys. Res. Lett. 1993, 20, 129–132. [Google Scholar] [CrossRef]
  61. 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]
  62. Andreae, M.O.; Crutzen, P.J. Atmospheric Aerosols: Biogeochemical Sources and Role in Atmospheric Chemistry. Science 1997, 276, 1052–1058. [Google Scholar] [CrossRef]
  63. Simmons, J.S.; Klemedtsson, L.; Hultberg, H.; Hines, M.E. Consumption of Atmospheric Carbonyl Sulfide by Coniferous Boreal Forest Soils. J. Geophys. Res. Atmos. 1999, 104, 11569–11576. [Google Scholar] [CrossRef]
  64. Mellillo, J.M.; Steudler, P.A. The Effect of Nitrogen Fertilization on the COS and CS2 Emissions from Temperature Forest Soils. J. Atmos. Chem. 1989, 9, 411–417. [Google Scholar] [CrossRef]
  65. Yi, Z.; Wang, X.; Sheng, G.; Zhang, D.; Zhou, G.; Fu, J. Soil Uptake of Carbonyl Sulfide in Subtropical Forests with Different Successional Stages in South China. J. Geophys. Res. 2007, 112, D08302. [Google Scholar] [CrossRef]
  66. Kitz, F.; Gerdel, K.; Hammerle, A.; Laterza, T.; Spielmann, F.M.; Wohlfahrt, G. In Situ Soil COS Exchange of a Temperate Mountain Grassland under Simulated Drought. Oecologia 2017, 183, 851–860. [Google Scholar] [CrossRef] [PubMed]
  67. Bunk, R.; Behrendt, T.; Yi, Z.; Andreae, M.O.; Kesselmeier, J. Exchange of Carbonyl Sulfide (OCS) between Soils and Atmosphere under Various CO2 Concentrations. J. Geophys. Res. Biogeosci. 2017, 122, 1343–1358. [Google Scholar] [CrossRef]
  68. Conrad, R. Soil Microorganisms as Controllers of Atmospheric Trace Gases (H2, CO, CH4, OCS, N2O, and NO). Microbiol. Rev. 1996, 60, 609–640. [Google Scholar] [CrossRef] [PubMed]
  69. 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]
  70. Kesselmeier, J.; Teusch, N.; Kuhn, U. Controlling Variables for the Uptake of Atmospheric Carbonyl Sulfide by Soil. J. Geophys. Res. Atmos. 1999, 104, 11577–11584. [Google Scholar] [CrossRef]
  71. Stimler, K.; Montzka, S.A.; Berry, J.A.; Rudich, Y.; Yakir, D. Relationships between Carbonyl Sulfide (COS) and CO2 during Leaf Gas Exchange. New Phytol. 2010, 186, 869–878. [Google Scholar] [CrossRef]
  72. Rennenberg, H. The Fate of Excess Sulfur in Higher Plants. Annu. Rev. Plant Physiol. 1984, 35, 121–153. [Google Scholar] [CrossRef]
  73. Protoschill-Krebs, G.; Kesselmeier, J. Enzymatic Pathways for the Consumption of Carbonyl Sulphide (COS) by Higher Plants. Bot. Acta 1992, 105, 206–212. [Google Scholar] [CrossRef]
  74. Matsumoto, H.; Hylin, J.W.; Miyahara, A. Methemoglobinemia in Rats Injected with 3-Nitropropanoic Acid, Sodium Nitrite, and Nitroethane. Toxicol. Appl. Pharmacol. 1961, 3, 493–499. [Google Scholar] [CrossRef]
  75. Lloyd, J.; Farquhar, G.D. 13C Discrimination during CO2 Assimilation by the Terrestrial Biosphere. Oecologia 1994, 99, 201–215. [Google Scholar] [CrossRef]
  76. Burnell, J.N.; Hatch, M.D. Low Bundle Sheath Carbonic Anhydrase Is Apparently Essential for Effective C4 Pathway Operation. Plant Physiol. 1988, 86, 1252–1256. [Google Scholar] [CrossRef] [PubMed]
  77. Davidson, C.; Amrani, A.; Angert, A. Carbonyl Sulfide Sulfur Isotope Fractionation During Uptake by C3 and C4 Plants. J. Geophys. Res. Biogeosci. 2022, 127, e2022JG007035. [Google Scholar] [CrossRef]
  78. Gillon, J.S.; Yakir, D. Naturally Low Carbonic Anhydrase Activity in C4 and C3 Plants Limits Discrimination against C18OO during Photosynthesis. Plant Cell Environ. 2000, 23, 903–915. [Google Scholar] [CrossRef]
  79. 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]
  80. Chengelis, C.P.; Neal, R.A. Studies of Carbonyl Sulfide Toxicity: Metabolism by Carbonic Anhydrase. Toxicol. Appl. Pharmacol. 1980, 55, 198–202. [Google Scholar] [CrossRef]
  81. Ingham, E.R.; Horton, K.A. Bacterial, Fungal and Protozoan Responses to Chloroform Fumigation in Stored Soil. Soil Biol. Biochem. 1987, 19, 545–550. [Google Scholar] [CrossRef]
  82. Blankinship, J.C.; Becerra, C.A.; Schaeffer, S.M.; Schimel, J.P. Separating Cellular Metabolism from Exoenzyme Activity in Soil Organic Matter Decomposition. Soil Biol. Biochem. 2014, 71, 68–75. [Google Scholar] [CrossRef]
  83. Maire, V.; Alvarez, G.; Colombet, J.; Comby, A.; Despinasse, R.; Dubreucq, E.; Joly, M.; Lehours, A.-C.; Perrier, V.; Shahzad, T.; et al. An Unknown Oxidative Metabolism Substantially Contributes to Soil CO2 Emissions. Biogeosciences 2013, 10, 1155–1167. [Google Scholar] [CrossRef]
  84. Saska, P.; Skuhrovec, J.; Tylová, E.; Platková, H.; Tuan, S.J.; Hsu, Y.T.; Vítámvás, P. Leaf structural traits rather than drought resistance determine aphid performance on spring wheat. J. Pest Sci. 2021, 94, 423–434. [Google Scholar] [CrossRef]
  85. 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]
  86. Chen, B.; Wang, P.; Wang, S.; Liu, Z.; Croft, H. Evaluation of Leaf-To-Canopy Upscaling Approaches for Simulating Canopy Carbonyl Sulfide Uptake and Gross Primary Productivity. J. Geophys. Res. Biogeosci. 2024, 129, e2023JG007521. [Google Scholar] [CrossRef]
  87. Caird, M.A.; Richards, J.H.; Donovan, L.A. Nighttime Stomatal Conductance and Transpiration in C3 and C4 Plants. Plant Physiol. 2007, 143, 4–10. [Google Scholar] [CrossRef] [PubMed]
  88. Costa, J.M.; Monnet, F.; Jannaud, D.; Leonhardt, N.; Ksas, B.; Reiter, I.M.; Pantin, F.; Genty, B. OPEN ALL NIGHT LONG: The Dark Side of Stomatal Control. Plant Physiol. 2015, 167, 289–294. [Google Scholar] [CrossRef]
  89. Liu, J.; Geng, C.; Mu, Y.; Zhang, Y.; Xu, Z.; Wu, H. Exchange of Carbonyl Sulfide (COS) between the Atmosphere and Various Soils in China. Biogeosciences 2010, 7, 753–762. [Google Scholar] [CrossRef]
  90. 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]
  91. Devai, I.; DeLaune, R.D. Trapping Efficiency of Various Solid Adsorbents for Sampling and Quantitative Gas Chromatographic Analysis of Carbonyl Sulfide. Anal. Lett. 1997, 30, 187–198. [Google Scholar] [CrossRef]
  92. Smith, N.A.; Kelly, D.P. Oxidation of Carbon Disulphide as the Sole Source of Energy for the Autotrophic Growth of Thiobacillus Thioparus Strain TK-m. Microbiology 1988, 134, 3041–3048. [Google Scholar] [CrossRef]
  93. Van Diest, H.; Kesselmeier, J. Soil Atmosphere Exchange of Carbonyl Sulfide (COS) Regulated by Diffusivity Depending on Water-Filled Pore Space. Biogeosciences 2008, 5, 475–483. [Google Scholar] [CrossRef]
  94. Drenovsky, R.E.; Vo, D.; Graham, K.J.; Scow, K.M. Soil Water Content and Organic Carbon Availability Are Major Determinants of Soil Microbial Community Composition. Microb. Ecol. 2004, 48, 424–430. [Google Scholar] [CrossRef] [PubMed]
  95. Schenk, S.; Kesselmeier, J.; Anders, E. How Does the Exchange of One Oxygen Atom with Sulfur Affect the Catalytic Cycle of Carbonic Anhydrase? Chem. Eur. J. 2004, 10, 3091–3105. [Google Scholar] [CrossRef]
  96. Kitz, F.; Spielmann, F.M.; Hammerle, A.; Kolle, O.; Migliavacca, M.; Moreno, G.; Ibrom, A.; Krasnov, D.; Noe, S.M.; Wohlfahrt, G. Soil COS Exchange: A Comparison of Three European Ecosystems. Glob. Biogeochem. Cycles 2020, 34, e2019GB006202. [Google Scholar] [CrossRef]
  97. Sauze, J.; Ogée, J.; Maron, P.-A.; Crouzet, O.; Nowak, V.; Wohl, S.; Kaisermann, A.; Jones, S.P.; Wingate, L. The Interaction of Soil Phototrophs and Fungi with pH and Their Impact on Soil CO2, CO18O and OCS Exchange. Soil Biol. Biochem. 2017, 115, 371–382. [Google Scholar] [CrossRef] [PubMed]
  98. Li, X.S.; Sato, T.; Ooiwa, Y.; Kusumi, A.; Gu, J.-D.; Katayama, Y. Oxidation of Elemental Sulfur by Fusarium Solani Strain THIF01 Harboring Endobacterium Bradyrhizobium sp. Microb. Ecol. 2010, 60, 96–104. [Google Scholar] [CrossRef]
  99. 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]
  100. 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]
  101. 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]
  102. Behrendt, T.; Catão, E.C.P.; Bunk, R.; Yi, Z.; Schweer, E.; Kolb, S.; Kesselmeier, J.; Trumbore, S. Microbial Community Responses Determine How Soil–Atmosphere Exchange of Carbonyl Sulfide, Carbon Monoxide, and Nitric Oxide Responds to Soil Moisture. Soil 2019, 5, 121–135. [Google Scholar] [CrossRef]
  103. Parazoo, N.C.; Barnes, E.; Worden, J.; Harper, A.B.; Bowman, K.B.; Frankenberg, C.; Wolf, S.; Litvak, M.; Keenan, T.F. Influence of ENSO and the NAO on Terrestrial Carbon Uptake in the Texas-Northern Mexico Region. Glob. Biogeochem. Cycles 2015, 29, 1247–1265. [Google Scholar] [CrossRef]
  104. Piao, S.; Sitch, S.; Ciais, P.; Friedlingstein, P.; Peylin, P.; Wang, X.; Ahlström, A.; Anav, A.; Canadell, J.G.; Cong, N.; et al. Evaluation of Terrestrial Carbon Cycle Models for Their Response to Climate Variability and to CO2 Trends. Glob. Chang. Biol. 2013, 19, 2117–2132. [Google Scholar] [CrossRef]
  105. McManus, J.B. Application of Quantum Cascade Lasers to High-Precision Atmospheric Trace Gas Measurements. Opt. Eng. 2010, 49, 111124. [Google Scholar] [CrossRef]
  106. 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, 2008GL034270. [Google Scholar] [CrossRef]
  107. Glatthor, N.; Höpfner, M.; Baker, I.T.; Berry, J.; Campbell, J.E.; Kawa, S.R.; Krysztofiak, G.; Leyser, A.; Sinnhuber, B.-M.; Stiller, G.P.; et al. Tropical Sources and Sinks of Carbonyl Sulfide Observed from Space. Geophys. Res. Lett. 2015, 42, 10082–10090. [Google Scholar] [CrossRef]
  108. Serio, C.; Montzka, S.A.; Masiello, G.; Carbone, V. Trend and Multi-Frequency Analysis Through Empirical Mode Decomposition: An Application to a 20-Year Record of Atmospheric Carbonyl Sulfide Measurements. J. Geophys. Res. Atmos. 2023, 128, e2022JD038207. [Google Scholar] [CrossRef]
  109. Remaud, M.; Chevallier, F.; Maignan, F.; Belviso, S.; Berchet, A.; Parouffe, A.; Abadie, C.; Bacour, C.; Lennartz, S.; Peylin, P. Plant Gross Primary Production, Plant Respiration and Carbonyl Sulfide Emissions over the Globe Inferred by Atmospheric Inverse Modelling. Atmos. Chem. Phys. 2022, 22, 2525–2552. [Google Scholar] [CrossRef]
  110. Haynes, K.D.; Baker, I.T.; Denning, A.S.; Stöckli, R.; Schaefer, K.; Lokupitiya, E.Y.; Haynes, J.M. Representing Grasslands Using Dynamic Prognostic Phenology Based on Biological Growth Stages: 1. Implementation in the Simple Biosphere Model (SiB4). J. Adv. Model. Earth Syst. 2019, 11, 4423–4439. [Google Scholar] [CrossRef]
  111. 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]
  112. Vesala, T.; Kohonen, K.-M.; Kooijmans, L.M.J.; Praplan, A.P.; Foltýnová, L.; Kolari, P.; Kulmala, M.; Bäck, J.; Nelson, D.; Yakir, D.; et al. Long-Term Fluxes of Carbonyl Sulfide and Their Seasonality and Interannual Variability in a Boreal Forest. Atmos. Chem. Phys. 2022, 22, 2569–2584. [Google Scholar] [CrossRef]
  113. Borrell, P.; Borrell, P.M. (Eds.) Transport and Chemical Transformation of Pollutants in the Troposphere: An Overview of the Work of EUROTRAC; Springer: Berlin/Heidelberg, Germany, 2000. [Google Scholar] [CrossRef]
  114. Taylor, G.E., Jr.; McLaughlin, S.B., Jr.; Shriner, D.S.; Selvidge, W.J. The Flux of Sulfur-Containing Gases to Vegetation. Atmos. Environ. (1967) 1983, 17, 789–796. [Google Scholar] [CrossRef]
  115. Spielmann, F.M.; Hammerle, A.; Kitz, F.; Gerdel, K.; Alberti, G.; Peressotti, A.; Delle Vedove, G.; Wohlfahrt, G. On the Variability of the Leaf Relative Uptake Rate of Carbonyl Sulfide Compared to Carbon Dioxide: Insights from a Paired Field Study with Two Soybean Varieties. Agric. For. Meteorol. 2023, 338, 109504. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of major COS sources, sinks, and fluxes in terrestrial ecosystems. Arrow up represents COS emission into the atmosphere, arrow down represents COS absorption from the atmosphere. ① represents COS transmission between the troposphere and the stratosphere. ② represents COS indirect transmission between vegetation and soil.
Figure 1. Schematic diagram of major COS sources, sinks, and fluxes in terrestrial ecosystems. Arrow up represents COS emission into the atmosphere, arrow down represents COS absorption from the atmosphere. ① represents COS transmission between the troposphere and the stratosphere. ② represents COS indirect transmission between vegetation and soil.
Atmosphere 15 00778 g001
Figure 2. Global average land COS uptake from observing and modeling studies. Negative values indicate sink. Year represents the results of research published in different years, specifically: 1993 (Chin et al., 1993 [37]); 2000 (Watts et al., 2000 [41]); 2007 (Montzka et al., 2007 [29]); 2008 (Suntharalingam et al., 2008 [38]); 2013 (Berry et al., 2013 [39]); 2015 (Launois et al., 2015 [40]); 2021 (Ma et al., 2021 [36]).
Figure 2. Global average land COS uptake from observing and modeling studies. Negative values indicate sink. Year represents the results of research published in different years, specifically: 1993 (Chin et al., 1993 [37]); 2000 (Watts et al., 2000 [41]); 2007 (Montzka et al., 2007 [29]); 2008 (Suntharalingam et al., 2008 [38]); 2013 (Berry et al., 2013 [39]); 2015 (Launois et al., 2015 [40]); 2021 (Ma et al., 2021 [36]).
Atmosphere 15 00778 g002
Figure 3. COS absorption and metabolism pathways in C3 (left) and C4 (right) plant leaves.
Figure 3. COS absorption and metabolism pathways in C3 (left) and C4 (right) plant leaves.
Atmosphere 15 00778 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, J.; Shen, L.; Zhang, Y.; Liu, Y.; Wu, J.; Wang, A. Carbonyl Sulfide (COS) in Terrestrial Ecosystem: What We Know and What We Do Not. Atmosphere 2024, 15, 778. https://doi.org/10.3390/atmos15070778

AMA Style

Li J, Shen L, Zhang Y, Liu Y, Wu J, Wang A. Carbonyl Sulfide (COS) in Terrestrial Ecosystem: What We Know and What We Do Not. Atmosphere. 2024; 15(7):778. https://doi.org/10.3390/atmos15070778

Chicago/Turabian Style

Li, Jiaxin, Lidu Shen, Yuan Zhang, Yage Liu, Jiabing Wu, and Anzhi Wang. 2024. "Carbonyl Sulfide (COS) in Terrestrial Ecosystem: What We Know and What We Do Not" Atmosphere 15, no. 7: 778. https://doi.org/10.3390/atmos15070778

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