Next Article in Journal
Suppression of NOX2-Derived Reactive Oxygen Species (ROS) Reduces Epithelial-to-MesEnchymal Transition Through Blocking SiO2-Regulated JNK Activation
Next Article in Special Issue
Optical Properties and Molecular Composition of Fine Organic Aerosols in Nanjing, China: A Comparison of 2019 and 2023
Previous Article in Journal
Immobilization of Inorganic Phosphorus on Soils by Zinc Oxide Engineered Nanoparticles
Previous Article in Special Issue
Atmospheric Heavy Metal Pollution Characteristics and Health Risk Assessment Across Various Type of Cities in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Biogenic Volatile Organic Compounds from Plants: Research Progress and Future Prospects

1
College of Resources and Environmental Engineering, Guizhou University, Guiyang 550025, China
2
College of Environmental Science and Engineering, Beijing Forestry University, Beijing 100083, China
3
State Key Laboratory of Environmental Criteria and Risk Assessment, Chinese Research Academy of Environmental Sciences, Beijing 100012, China
4
Guizhou Research and Designing Institute of Environmental Sciences, Guizhou Academy of Environmental Science and Design, Guiyang 550081, China
5
North Alabama International College of Engineering and Technology, Guizhou University, Guiyang 550025, China
6
Department of Fisheries, Faculty of Biological Sciences, University of Dhaka, Dhaka 1000, Bangladesh
7
Key Laboratory of Karst Georesources and Environment, Ministry of Education, College of Resources and Environmental Engineering, Guizhou University, Guiyang 550025, China
*
Authors to whom correspondence should be addressed.
Toxics 2025, 13(5), 364; https://doi.org/10.3390/toxics13050364
Submission received: 4 February 2025 / Revised: 25 April 2025 / Accepted: 28 April 2025 / Published: 30 April 2025
(This article belongs to the Special Issue Source and Components Analysis of Aerosols in Air Pollution)

Abstract

:
Biogenic volatile organic compounds (BVOCs) emitted by plants contribute to secondary air pollution through photochemical reactions in sunlight. Due to the influence of multiple factors, accurately characterizing and quantifying the emission of BVOCs from plant sources is challenging, which poses significant obstacles to the effective management and control of BVOCs. Therefore, this paper summarizes the emission mechanisms of BVOCs from plants, explores the primary factors influencing variations in the emission rates of these compounds, and evaluates the advantages and limitations of contemporary “measurement-modeling” methods for characterizing BVOC emissions. It is concluded that current measurement techniques still need to be further developed to meet the criteria of simplicity, affordability, and high precision simultaneously, and in terms of modeling and prediction studies, there is a lack of in-depth research on the atmospheric chemistry of BVOCs and the synergistic effects of multiple factors. Finally, it is suggested to leverage interdisciplinary strengths to develop advanced measurement technologies and high-resolution models for monitoring volatile compounds. Additionally, strategically selecting low-BVOC tree species in pollution-vulnerable urban areas—contingent on rigorous ecological assessments—combined with stringent controls on anthropogenic precursors (e.g., anthropogenic volatile organic compounds (AVOCs)) could serve as a complementary measure to mitigate secondary pollution.

Graphical Abstract

1. Introduction

As public awareness of environmental protection increases, attention to environmental issues has intensified, particularly concerning air pollution, which directly affects human health. Among various air pollutants, volatile organic compounds (VOCs) stand out not only for their potential toxicity but also for their propensity to react with other pollutants, leading to secondary pollution. In 1990, the U.S. enacted the Clean Air Act Amendments, which listed 189 hazardous air pollutants (HAPs), including anthropogenic volatile organic compounds (AVOCs) such as benzene and toluene. The subsequent European Union Industrial Emissions Directive (2010/75/EU) also stipulates that industries emitting VOCs must implement Best Available Technology (BAT) to achieve a 60–90% reduction rate. And China also began to regulate VOCs at the beginning of the 21st century with the promulgation of the Volatile Organic Compounds (VOCs) Emission Control Standard. However, significant challenges remain for VOCs in non-regulated areas and emerging emission sources. Notably, biogenic volatile organic compounds (BVOCs) account for 90% of global VOC emissions [1,2,3]. The primary source of global BVOCs is plants, which naturally emit these compounds. However, their photochemical derivatives, like ozone (O3) and secondary organic aerosols (SOAs), may pose environmental and health risks in regions with elevated anthropogenic influences [4]. Although emissions from individual plants are relatively minor, forests, grasslands, wetlands, and agricultural lands—covering 31% of the Earth’s land surface—substantially amplify BVOC emissions [5,6,7]. Therefore, understanding the emission mechanisms of BVOCs and their synergistic effects with anthropogenic pollutants is critical for developing integrated pollution prevention and control strategies.
BVOCs are the main precursors of ozone (O3), accounting for 20% of total O3 production [8]. In addition, due to the photosynthetic characteristics of plants, the photochemical reaction activity of plant-generated BVOCs is significantly higher than that of anthropogenic volatile organic compounds (AVOCs) [9,10]. Under sunlight, BVOCs readily undergo photochemical reactions, contributing to the formation of biogenic secondary organic aerosols (BSOAs), which play a significant role in the global secondary organic aerosol (SOA) budget [11,12,13,14]. Although BVOCs are naturally emitted and not directly anthropogenic, their photochemical derivatives—tropospheric ozone (O3) and secondary organic aerosol (SOA)—are recognized as hazardous pollutants under elevated anthropogenic influences. Furthermore, a comprehensive understanding of BVOCs is vital for advancing insights into Earth’s ecological and atmospheric systems.
The historical evolution of BVOCs reveals significant progress driven by advancements in mechanistic studies and measurement techniques. In the 1950s, Sanadze [15] first observed the emission of gaseous organic substances from plants, initially identifying isoprene as a key compound [16]. With the application of gas chromatography, it became evident that BVOCs include a diverse array of around 30,000 hydrocarbons and oxides, including isoprene (C5H8), monoterpenes (C10H16), sesquiterpenes (C15H24), and other BVOCs [17,18,19]. These compounds exhibit diverse chemical structures and properties, necessitating specialized detection methods for their identification and quantification. By gas chromatography–mass spectrometry (GC-MS) technology, Kesselmeier and Staudt [20] provided insights into BVOCs biosynthesis, emission inventories, and the relationship between emissions, plant physiology, temperature, radiation, and ecophysiological functions. Helmig et al. [21] further advanced BVOCs detection by employing thermal desorption–gas chromatography–mass spectrometry (TDS-GC-MS), which enabled the analysis of a broader range of polar compounds in BVOCs and facilitated deeper mechanistic studies of BVOCs.
Despite these advances, challenges still persist in detecting and identifying the types of BVOCs from plants. Atmospheric concentrations of BVOCs from plants are typically very low, and their high reactivity leads to rapid chemical transformation or adsorption within sampling equipment, complicating accurate measurement. Additionally, temporal and spatial variations in BVOC emissions—driven by factors such as industrial activity, regional environmental differences, and seasonal climate changes—introduce further complexities [22]. On larger scales, studies by Papiez et al. [23] and Kramshøj et al. [24] revealed that cumulative emissions of isoprene, monoterpenes, and other BVOCs, amplified by anthropogenic climate perturbations, contribute to atmospheric processes (e.g., O3 formation) and carbon–climate feedback, thereby indirectly influencing climate warming trends.
Recent advancements in computational methodologies have revolutionized the study of BVOC emissions from vegetation. Research paradigms have transitioned from foundational investigations into biochemical mechanisms and empirical sampling analyses to the incorporation of advanced machine learning architectures. These computational innovations facilitate the predictive modeling of BVOCs’ multifactorial impacts, including climate adaptation dynamics, O3 generation pathways, and SOA formation kinetics. Guenther et al. [3,25,26] pioneered BVOC emissions modeling, developing global models that laid the foundation for quantifying BVOC emissions. On this basis, Pierce [27] studied the aerosol impacts on the climate systems by the improved atmospheric chemistry models. More recently, Ng et al. [28] explored interactions between nitrate radicals (•NO3) and BVOCs during SOA formation and highlighted the importance of such interactions in improving atmospheric model predictions. Tang et al. [29] combined dynamic vegetation models, such as LPJ-GUESS v4.1, with global chemical transport models like TM5-MP, demonstrating that the spatial pattern of the BVOCs-driven SOA impact was strongly influenced by climate-induced vegetation changes.
Now, the research on BVOCs is no longer limited to a single discipline. Studies by Lun et al. [30] and Yang et al. [31] have contributed significantly to initiating BVOCs research by reviewing sampling and analysis methods, emission inventories, and emission factors. As disciplines such as chemistry, biology, mathematics, and computer science intersect with environmental science, new perspectives in BVOCs from plant research are emerging. This paper reviews recent studies on BVOCs emission sources, chemical composition, and reaction mechanisms; discusses their environmental impact factors; and evaluates effective mitigation strategies. Additionally, it highlights future interdisciplinary research trends, providing valuable insights for addressing BVOCs-related atmospheric pollution.

2. Methodology

To comprehensively review and summarize the research progress of BVOCs, this study adopted an interdisciplinary narrative review framework (Figure 1). A systematic literature search was conducted across multiple databases, including Web of Science, Elsevier ScienceDirect, and the China National Knowledge Infrastructure (CNKI). Keywords such as “BVOCs”, “BVOCs emission”, and “atmospheric oxidation” were employed for thematic searches, yielding an initial collection of 2483 articles. Following the screening and analysis of search results, duplicate records were removed, resulting in 2152 articles selected for detailed evaluation. Subsequently, 654 articles unrelated to the research scope were excluded based on titles and abstracts. Finally, through full-text precision assessments, 991 articles were further excluded due to data redundancy, incomplete methodological descriptions, or lack of environmental mechanism relevance, ultimately retaining 507 research articles for in-depth analysis.
As shown in Figure 2, the reviewed literature comprises 132 domestic Chinese articles and 375 international articles, spanning from 1999 to 2024. The earliest relevant articles originated in the United States, followed by contributions from South Korea. Between 1999 and 2009, only 28 articles were published. Since 2010, however, the number of publications has steadily increased, reaching an explosion in 2020 and reaching its peak in 2024. This trend highlights the growing global interest in BVOCs research, particularly in understanding their atmospheric interactions and environmental impacts.

3. BVOC Emissions Characterization Methods

3.1. Methods of Emission Measurement

3.1.1. Single-Plant Scale Sampling Methods

To comprehensively understand the impact of BVOC emissions on atmospheric chemistry and the environment, it is essential to accurately monitor their emission flux. Currently, emission measurements of BVOCs from a single plant focus on the leaf scale, and the main methods are divided into on-line and off-line techniques (Table 1), with each category encompassing various measurement techniques and technologies, which involve analytical techniques as shown in Table 2.
(1)
On-line measurement
On-line measurement methods are frequently employed for monitoring environmental conditions in the field. On-line detection can analyze the emission concentration of BVOCs in the atmosphere and their temporal variations in real time. This method typically involves direct air sampling and immediate analysis using advanced instruments [55,56,57].
Representative on-line measurement methods include proton transfer reaction–mass spectrometry (PTR-MS) [58,59], Fourier transform infrared spectroscopy (FTIR) [60,61], portable gas chromatography–flame ionization detector (GC-FID), and portable GC-MS [62,63]. PTR-MS is particularly useful for trace gas analysis and is widely used in outdoor atmospheric measurements. To study the response of temperate deciduous tree species to flooding stress, Copolovici et al. [64] used proton transfer reaction–mass spectrometer (PTR-MS, high sensitivity version) to measure VOC concentration. Gómez et al. [34] routinely determined VOCs and qualitatively identified monoterpene compounds in ambient air using a gas chromatography system with GC-FID. You et al. [65] upgraded the traditional portable gas chromatograph and used it for continuous and on-site determination of trace volatile organic compounds, especially aromatic volatile organic compounds, in air samples.
The FTIR device can identify and quantify multiple BVOCs in air by measuring the infrared absorption spectrum of the gas without sample pretreatment and can simultaneously detect multiple compounds. Kalalian et al. [35] used FTIR and solid-phase microextraction sampling combined with a gas chromatography–mass spectrometry (SPME-GC/MS) in situ analysis of the reaction products of O3 and three different BVOCs to predict the formation of photooxidants such as SOA. In order to check the health of plants, Sharma et al. [66] developed and applied a fully automated portable gas phase color spectrum for a rapid and in situ analysis of BVOCs from plants. Cicciol et al. [67] measured the vertical fluxes of volatile organic compounds from vegetation in a Mediterranean forested area using PTR-MS.
In addition to these methods, electronic noses (E-noses) represent an emerging technology for real-time environmental monitoring. E-noses offer high sensitivity and rapid response and eliminate the need for the complex pretreatment processes associated with gas chromatography sensors [36,68]. Feng et al. [69] used E-noses, headspace solid-phase microextraction–gas chromatography–mass spectrometry, and headspace–gas chromatography–ion mobility spectrometry to identify and characterize volatile organic compounds in eight kinds of Zanthoxylum bungeanum with geographical indications in China, providing a useful basis for the study of the differences in floral aroma. However, while E-noses are advantageous for qualitative and semi-quantitative analysis, the high complexity of environmental backgrounds can complicate the practical application of the E-noses, and their analytical accuracy and repeatability remain limited.
(2)
Off-line measurement
Off-line measurement methods involve air sample collection and laboratory analysis. These methods are advantageous for their simplicity, high accuracy, and high sensitivity, but they lack real-time data.
(1)
Sample collection
Off-line measurement methods separate sampling from analysis, where samples are collected on-site and then transported to a laboratory for analysis. Sample preservation methods for the different biogenic volatile organic compound (BVOC) collection containers are shown in Table 3. These methods are simple to operate, highly accurate, and sensitive. However, because samples need to be transported, they cannot reflect real-time data.
Typical collection containers for BVOCs include sampling bags (such as Tedlar® Bags and Teflon Bags), adsorption tubes (such as Tenax® Tubes and Activated Charcoal Tubes), and sampling canisters (such as SUMMA Canisters, Stainless Steel Canisters, and Silonite Canisters). Of these, the most commonly used containers for ambient air sampling are SUMMA Canisters, which have special treatments such as silanization or fused silica coatings that minimize the adsorption or reaction of VOCs on the walls of the canisters, thus maintaining the chemical stability of the sample. Sampling methods mainly include static closed sampling and dynamic headspace sampling. Static closed sampling, a classic method for calculating gas emission rates, involves measuring the concentration of BVOCs in a chamber or bag containing the plant species over time to determine the emission rate. Lin et al. [70] demonstrated that static enclosure techniques and in situ sampling can accurately capture BVOC emissions from entire leaves, which can then be estimated using remote sensing data. To address the limitations of static enclosure techniques, a semi-static method has been developed. Li et al. [71] conducted repeated experiments to evaluate the performance of the semi-static method for measuring BVOC emissions, confirming the feasibility of semi-static closed techniques as a screening tool for characterizing monoterpene emissions.
While static methods allow for rapid measurements and lower detection limits, they are susceptible to environmental conditions; factors such as temperature, pressure, and humidity can easily alter the equilibrium of BVOCs between the sample and the gas phase [72]. To address these limitations and meet the measurement requirements of various different scenarios, researchers have developed a dynamic headspace sampling method. This dynamic system features air circulation capabilities, effectively reducing discrepancies in environmental parameters inside and outside the system, enabling comprehensive environmental control for accurate BVOCs emission measurements. Vedel-Petersen et al. [73] utilized dynamic enclosure techniques to collect BVOC emissions from Arctic plants; this approach reduced environmental influence and allowed for the study of emission characteristics and seasonal variations in BVOCs. Jing et al. [74] also used dynamic headspace sampling to investigate BVOC emissions in northern China, concluding that different tree species emit different components at different rates, further demonstrating the advantages of dynamic headspace sampling. Dynamic headspace sampling is the most commonly used method for measuring BVOC emissions from living plants.
(2)
Sample analysis
After field samples are collected, they are preserved and transported to the laboratory using specific methods to maintain sample integrity, as shown in Table 3. Once in the laboratory, these samples must typically be analyzed using various detection instruments within a specific timeframe to ensure accurate results.
At present, the most widely used separation and identification instrument is GC-MS; TDS-GC-MS is often used in the qualitative analysis study of BVOC emissions. Yuan et al. [75] quantified isoprene emissions from young hybrid poplar trees using TDS-GC-MS, and for the first time, investigated the combined effects of O3 and drought on isoprene emissions. In order to detect the main BVOC species emitted by Mediterranean plant species, Fattobene et al. [38] employed thermal desorption gas chromatography–mass spectroscopy (TD-GC-MS) to analyze the air on-site sampled in the field.
Based on GC-MS analysis methods, researchers have also improved and developed automatic thermal desorption–gas chromatography–mass spectrometry (TCT-GC-MS), thermal desorption gas chromatography–time-of-flight mass spectrometry (TD-GC/TOF-MS) [40], etc.
Table 3. Sample preservation methods for different biogenic volatile organic compound (BVOC) collection containers.
Table 3. Sample preservation methods for different biogenic volatile organic compound (BVOC) collection containers.
Collection ContainersPretreatment RequirementsStorage ConditionsMaximum Preservation TimeApplicable Compound RangeRef.
Tedlar BagsNitrogen flushed 3 times, light-proof aluminum foil wrapped4 °C, <30% humidity48 hNonpolar terpenes (C5-C10) [76,77]
Tenax TubesAging (320 °C, 2 h), vacuum sealed in aluminum foil pouch−20 °C with desiccant7–30 daysSemi-volatile VOCs (C10-C15) [78,79]
Activated Charcoal TubesAged 1 h at 400 °C, sealed in moisture-proof container4 °C, avoid condensation14 daysHighly volatile VOCs (C2-C8) [80,81]
SUMMA CanistersVacuum baked at 160 °C for 12 h, nitrogen filled to 30 psiProtected from light at room temperature, pressure monitoring 30 daysBroad spectrum VOCs (C2-C15) [77,82]
Typical collection containers for BVOCs include sampling bags (such as Tedlar® Bags and Teflon Bags), adsorption tubes (such as Tenax® Tubes and Activated Charcoal Tubes), and sampling canisters (such as SUMMA Canisters, Stainless Steel Canisters, and Silonite Canisters).

3.1.2. Canopy Scale Flux Measurement Methods

The emissions of BVOCs exhibit high spatiotemporal heterogeneity. Traditional leaf-scale studies are constrained by the limited spatial representativeness of sampling methods. However, canopy-scale flux measurement methods such as Eddy Covariance (EC) and Relaxed Eddy Accumulation (REA), which directly capture integrated canopy-level fluxes, significantly enhance spatial representativeness, temporal continuity, and data comprehensiveness [83,84]. While early EC technology provides direct turbulent flux data, its reliability for BVOCs—due to their ultra-low concentrations and adsorption losses—requires advanced sensors like PTR-MS [85]. For example, Spirig et al. [86] combined EC and PTR-MS to measure BVOCs canopy fluxes in a mixed deciduous forest in northwestern Germany, validating EC’s reliability under complex heterogeneous canopy conditions, with isoprene and monoterpenes identified as dominant emissions. Misztal et al. [87] employed EC-PTR-MS to directly quantify BVOC fluxes in Southeast Asian oil palm ecosystems, revealing isoprene’s predominant role (84% of midday fluxes) and significant contributions from floral emissions. To enhance signal-to-noise ratios, REA integrates airflow separation and offline analysis (e.g., GC-MS) for trace BVOC detection. Bai et al. [88] applied REA and gradient techniques to conduct long-term continuous canopy flux measurements of BVOCs in a subtropical pine plantation, uncovering pronounced diurnal, seasonal, and interannual variations in isoprene and monoterpene emissions.
Subsequent advancements include Disjunct Eddy Covariance (DEC), which combines EC’s turbulence-phase data with REA’s intermittent sampling, broadening EC’s applicability [89,90,91]. Emerging methods like Disjunct Eddy Accumulation (DEA) leverage ultra-low-frequency sampling and automated accumulation to reduce power consumption and costs, enabling long-term field observations [92]. Current trends focus on integrating multiple methods and utilizing Unmanned Aerial Vehicle (UAV) platforms to expand 3D flux monitoring capabilities, thereby comprehensively elucidating the spatiotemporal patterns of BVOC emissions and their impacts on atmospheric chemistry [93,94,95].

3.2. Model-Based Estimation of BVOCs Emission

Although canopy-scale flux measurement techniques provide high-precision real-time flux data through direct observation, effectively reflecting plant responses to environmental changes, this method has limitations in spatial coverage and long-term monitoring, making it unable to comprehensively capture the dynamic variations in BVOCs.
Emission models bridge critical knowledge gaps by synthesizing empirical observations with environmental drivers (e.g., meteorology and land use patterns), enabling dynamic predictions of BVOC fluxes across heterogeneous ecosystems and under varying climatic scenarios. These tools are particularly vital for scaling plot-level measurements to regional or global assessments, where direct monitoring becomes logistically impractical. The development of BVOC emission models serves a dual purpose: (1) to extrapolate mechanistic insights from laboratory and field studies into predictive algorithms, and (2) to inform policy-relevant scenarios such as air quality management and climate mitigation strategies.
Thus, in the study of BVOC emissions in a certain area and at a large scale, the model-based estimation method is mainly used [96,97]. The model simulation method is based on remote sensing data to obtain the regional leaf area index and leaf biomass, combined with emission factors, meteorological data, and other information to use model simulation to obtain the total emission of plant-derived volatile organic pollutants, including the Guenther model (Guenther), model of emissions of gases and aerosol from nature (MEGAN), biological source emission inventory system model (BEIS), and so on. Their main characteristics and applicable scenarios are summarized in Figure 3. In recent years, the various models have been continuously upgraded and developed and are widely used in regional and global BVOC emission research.

3.2.1. Guenther

The isoprene emission model-G91, constructed by Guenther et al. [26] based on the photosynthetic electron transport model developed by Farquhar and the physiological characteristics of plants, provides an important framework for estimating inventories of biogenic non-methane hydrocarbon emission rates. In order to simulate the effects of light and leaf temperature on the emission of plant isoprene and monoterpenes, Guenther et al. [25] developed G93 on the basis of G91. Subsequently, based on the G91 and G93 algorithms, Guenther et al. [3] developed a global natural VOCs emission model, G95, based on relevant data, which has a high-resolution spatial grid to estimate hourly averaged emissions of organic compounds from natural sources. Chi et al. [98] used the Guenther model to convert vegetation volume and yield data, such as the forest and grassland resource inventory and vegetation maps, into leaf biomass data to estimate the space–time and source distribution of the emissions of BVOCs in China, and established a BVOC emissions inventory in mainland China. In order to explore the emission of BVOCs in Qinghai Province and its impact on air quality, Wang et al. [99] used the G95 method to estimate the total emissions of isoprene, monoterpenes, and other VOCs emitted by vegetation in Qinghai Province in 2021.

3.2.2. BEIS

Pierce et al. [27] established a biological emission inventory system model-BEIS1 based on the data of leaf biomass, land use, emission factors, and meteorological parameters obtained by observation or calculation to estimate VOCs emissions in vegetation and NO emissions in soil. Subsequently, Pierce et al. [100] developed the second version of the bio-emission inventory system BEIS2 on the basis of BEIS1. BEIS2 updated the land use inventory, emission factors, and environmental correction formulas following the G95 algorithm. Considering the factors such as plant growth environment, the National Center for Atmospheric Research (NCAR) and the United States Environmental Protection Agency (EPA) developed the global biosphere emission and interaction system—GloBEIS [30]. With the development of geographic information system technology, researchers have successively updated and developed BEIS3, BEIS4, etc. Feldman and Zheng et al. [101,102] used vegetation data and the GloBEIS model to estimate the total emission of plant-derived volatile organic compounds in southeastern Texas of the United States and in the Pearl River Delta of China.

3.2.3. MEGAN

Based on the G95 model and multi-source information such as satellite and ground observation data, Guenther et al. [2] constructed the natural gas and aerosol emission model, MEGAN, to estimate global terrestrial isoprene emissions. This model is suitable for regional- and global-scale chemical and transport modeling. With the deepening of research, researchers have successively developed MEGANv2.02 [103], MEGANv2.1 [104], MEGANv3.0 [105], and so on.
Sindelarova et al. [106] used the MEGANv2.1 model and MERRA meteorological field to establish a monthly global BVOC emissions dataset from 1980 to 2010. The model estimated the average annual BVOC emissions to be 760 Tg (C) yr−1. Using the improved MEGAN-BEIS361 model combination based on the MEGAN model, Wang et al. [107] simulated and predicted isoprene emissions and environmental concentrations in the United States. Jiang et al. [108] simulated biogenic isoprene emissions using the MEGAN3 response to drought and its impact on global atmospheric chemistry. Compared to the Guenther model and BEIS models, the MEGAN model is more widely used and is one of the commonly used models.
The operation of these three models requires various types of remote sensing data, meteorological data, and emission factors based on laboratory or field observations. They are widely used to obtain BVOC emissions across different temporal and spatial scales, providing effective methods for large-scale BVOC research. In recent years, building upon these three major models, researchers have successively developed new models such as the empirical model of BVOC emissions (EMBE) [109] and the Plant Specific Emission Model (PSEM) [67], offering new approaches for BVOC emission studies.

4. Emission Mechanisms of BVOCs from Plants

4.1. Emission Sources

The sources of BVOCs are highly diverse, both in urban and rural environments. In urban areas, both natural vegetation remnants (e.g., preserved woodlands) and cultivated green infrastructure (e.g., parks and street trees) serve as dominant sources of BVOCs. Notably, in certain cities, the spatial coverage of landscaped vegetation surpasses that of natural ecosystems, which may generate a distinct urban BVOCs signature compared to natural landscapes. Furthermore, the synergistic effects of environmental stressors—particularly the urban heat island effect and industrial pollutants—alter the compositional characteristics (e.g., shifts in isoprene-to-monoterpene ratios) and elevate the emission fluxes of plant-derived BVOCs, thereby intensifying their role in urban atmospheric chemistry [110,111]. Jing et al. [74] investigated BVOC emissions from dominant forest vegetation species in Beijing, finding that urban forest trees primarily emit isoprene and monoterpenes. Bao et al. [112] estimated BVOC emissions from 23 urban plants in Shenzhen and discovered that BVOC emissions from urban green spaces account for approximately 20.07% of the total BVOCs in the study region, with their estimated emissions in built-up areas (1.65 kt) nearly doubling conventional estimates (0.86 kt). This result challenges the historical marginalization of urban vegetation in BVOC emission inventories.
In rural areas, natural forest vegetation, agricultural crops, and wetlands are primary BVOC sources. Forests, in particular, are the largest contributors, as their emissions are highly influenced by natural environmental conditions [104,113,114]. Smiatek and Steinbrecher [115] studied the temporal and spatial variation in VOC emissions from German forests and found that the average annual BVOC emissions from forests during 1994–2003 were 366 Gg. Liu et al. [116] identified isoprene and monoterpenes as the primary emissions from pine and birch forests, with these compounds presenting the highest BVOC mixing ratios in forest ecosystems. Bamboo forests are also a major BVOC source [117]. Okumura et al. [118] found that 86% of bamboo species emit isoprene at rates ranging from 0.7 to 99.1 nmol m−2 s−1, with Phyllostachys glauca, Semiarundinaria fastuosa, and Bambusa pervariabilis being prominent sources.
Wetland plants, such as Phragmites australis, Typha angustifolia, Iris pseudacorus, and Cyperus rotundus, also produce BVOCs [119,120]. In agricultural landscapes, crops such as wheat, corn, and canola can release BVOCs, with corn and canola emitting methanol and acetaldehyde [121,122]. Additionally, microbial activity in soil produces small amounts of BVOCs, which, while minor in magnitude, are significant on a global scale [123,124,125,126]. Wester-Larsen et al. [127] reported variation in soil BVOC (SBVOC) concentrations based on different vegetation types, with areas of high vegetation cover and biodiversity, such as the Amazon rainforest and tropical forests in Southeast Asia, exhibiting higher BVOC emissions [128,129,130].
In both urban and rural areas, human activities have solidified the plant species in those regions (landscape plants in urban areas and economic crops in rural areas). If the emission sources of BVOCs in human-affected areas are not taken into account and higher precision emission estimates are not employed, there will be significant deviations in the understanding of BVOC emission characteristics. Therefore, studies on BVOC emissions require regional representation, with global research necessitating multi-regional composite studies that account for vegetation cover and topography. This paper primarily focuses on plant-derived BVOC emissions.

4.2. Compositions

BVOCs are broadly categorized into four main groups: isoprene, monoterpenes, sesquiterpenes, and oxygenated biogenic volatile organic compounds (OBVOCs), which include alcohols, aldehydes, ketones, and esters (Table 4). Among these, isoprenes and monoterpenes dominate, accounting for over 70% of the total annual BVOC emissions, particularly during the summer in temperate forests [131]. Liu et al. [132] revealed that isoprene emissions from Ficus virens could account for up to 85% of the total plant emissions. Similarly, monoterpene from Camphora officinarum contributes over 75% to total emissions and is highly reactive in forming O3 and SOA. To further illustrate the variability in BVOC emissions and concentrations, we present Table 5, which provides a comparative analysis of the mean concentrations of key BVOC components, such as isoprene and monoterpenes, across different study areas and climatic conditions [133].
Although sesquiterpenes and oxygenated biogenic volatile organic compounds (OBVOCs) are emitted in smaller quantities, they still significantly influence atmospheric chemistry. Liu et al. [134] confirmed that 2-methyl-3-buten-2-ol and methanol from pine forests undergo rapid oxidation, leading to SOA formation. While sesquiterpenes are emitted in lower amounts compared to isoprene and monoterpenes, they dominate in specific plant species such as Empetrum hermaphroditum and Cassiope tetragona [135]. Sesquiterpenes exhibit high oxidation reactivity with hydroxyl radicals (OH) and ozone (O3), resulting in rapid degradation and short atmospheric residence times (minutes to hours). Nevertheless, their oxidation products are important precursors of secondary organic aerosols (SOAs) [136,137,138]. Similarly, OBVOCs, such as esters and ketones, have short atmospheric lifetimes. For instance, certain ketones (e.g., methyl vinyl ketone) exhibit lifetimes of less than 1 h, where their low persistence is governed more by photolysis than by “lower chemical reactivity”. This limits their direct contribution to ozone formation but may enhance nighttime nitrate radical (NO3) chemistry [139,140,141,142]. The rapid dissociation of various BVOCs, including isoprene [143], α-pinene [144], β-pinene [145], d-limonene [146], and linalool [147], plays a critical role in O3 and SOA formation.
Table 4. Major categories of biogenic volatile organic compounds (BVOCs) and their global average emission values (notes: Ref., references).
Table 4. Major categories of biogenic volatile organic compounds (BVOCs) and their global average emission values (notes: Ref., references).
BVOCs CategoriesSynthesis MethodMain Emission Plants CategoryContribution RateGlobal c
(Tg C yr−1)
Ref.
1. IsoprenoidsMEPBroadleaf trees40–60%400–600[148,149,150]
2. MonoterpenesMVAConiferous trees25–40%40–180[151,152]
3. SesquiterpenesMEP and MVATropical rainforest plants
Herbaceous plants
5–15%20–40[153,154,155]
4. OBVOCsLOX/MEP/MVAHerbaceous plants;
trees; grasses
crops; flowers
10–25%–280[156,157]
Table 5. Mean concentrations of key biogenic volatile organic compounds (BVOCs) in different study areas and climatic conditions (Ref: references; -: no data).
Table 5. Mean concentrations of key biogenic volatile organic compounds (BVOCs) in different study areas and climatic conditions (Ref: references; -: no data).
Study AreaSeasonal Mean Concentrations (μg/m3)Annual Mean Concentrations (μg/m3)Ref.
SpringSummerAutumnWinter
Isoprene
Urban area of Beijing1.83 2.82 1.48 0.43 1.66 [133]
Suburban areas of Tianjin--1.61-- [158]
Urban area of Shenzhen2.074.261.09-1.06- [159]
Village of Guangzhou6.69-4.260.403.41 [160]
Urban area of Shanghai----0.10 [161]
Urban area of Houston----1.92 [162]
Urban area of Lille-0.88-0.330.58 [162]
Zurich-0.49-0.24- [162]
α-Pinene
Typical urban area of Beijing0.050.11 0.18 0.04 0.10 [133]
Mountainous areas of Sichuan----0.30 [163]
Guangzhou Forest Park-1.12-- [164]
β-Pinene
Typical urban area of Beijing0.05 0.003 0.006 0.0020.02 [133]
Mountainous areas of Sichuan----0.30 [163]

4.3. Synthetic Pathways

Biogenic volatile organic compounds (BVOCs) in plants originate from three principal biosynthetic routes: the methylerythritol phosphate (MEP) pathway localized in plastids, the mevalonate (MVA) pathway operating in cytoplasmic compartments, and the lipoxygenase (LOX) pathway associated with peroxisomal membranes [165,166,167,168]. The MEP pathway predominates in higher plant terpenoid biosynthesis, generating monoterpenoids and diterpenoids via sequential condensation of isopentenyl diphosphate (IPP) and dimethylallyl diphosphate (DMAPP) precursors synthesized from glyceraldehyde-3-phosphate and pyruvate [169,170,171,172,173]. Counterintuitive to previous assumptions, the MVA pathway—traditionally associated with fungal sesquiterpenogenesis—exhibits functional conservation in plant cytoplasmic metabolism [174,175]. This pathway converts acetyl-CoA through enzymatic cascades to yield IPP/DMAPP, primarily fueling sesquiterpene and phytosterol biosynthesis, with emerging evidence confirming its collaborative role with the MEP pathway in generating specific monoterpenes under stress conditions [176,177]. The LOX pathway initiates the oxidative cleavage of polyunsaturated fatty acids (e.g., linolenic acid), producing jasmonate phytohormones and green leaf volatiles that orchestrate systemic defense responses against biotic stressors [178,179,180,181]. Therefore, studies on the components and formation pathways of BVOCs should be analyzed on a case-by-case basis, and the synergistic effects of multiple factors may result in emission scenarios that are quite different from a single analysis.

5. Influencing Factors of BVOC Emissions

The release of BVOCs is a complex, multifactor process influenced by both intrinsic plant factors (plant types, physiological differences, and biotic stresses) and environmental factors (temperature, light, humidity, CO2 concentration, and O3 stress), which may affect the release rate and release characteristics of BVOCs in plants [182]. These factors and their influence on BVOC emissions are shown in Figure 4.

5.1. Intrinsic Plant Factors Influencing BVOC Emissions

5.1.1. Plant Types

Plant species are a key determinant of BVOC synthesis and release due to the differences in genetic control mechanisms and metabolic pathways, and the composition and release rate of BVOCs vary from different vegetation types [99]. Coniferous forests and broadleaf forests are two typical plant types, and their emission characteristics of BVOCs are different [183,184]. Coniferous forests mainly release monoterpenes and sesquiterpenes that may have more significant impacts on SOAs and the global climate system [185,186], while broadleaf forests mainly release short-chain BVOCs such as isoprene that are highly volatile and mainly affect photochemical smog and local O3 concentrations [187,188,189].

5.1.2. Physiological Differences

The physiological state and development stage of plants also play a critical role in BVOC emissions. Younger trees generally emit BVOCs at higher rates, while mature trees emit higher absolute quantities of these compounds. Schade et al. [190] and Street et al. [191] found that young pine branches and leaves release BVOCs at a rate 2–3 times higher than mature branches and leaves. Kuzma and Fall [192] and Rasulov et al. [193] observed lower isoprene emissions from young leaves to mature leaves both in Vicia faba and Populus alba, respectively.
Stomatal conductance is another critical physiological trait that influences BVOCs’ synthesis and emission. As the primary pathway for BVOC emission, stomatal conductance affects the partitioning of BVOCs between the gas and liquid phases within leaves’ release [194,195,196,197]. Leaf structure also plays a role in modulating the effect of stomata on VOC emission [198,199].

5.1.3. Biotic Stresses

Biotic stress, particularly pathogen infection and herbivore activity, can elevate BVOC emissions by 2–10-fold compared to unstressed plants, with emission profiles diverging based on stressor type (e.g., pathogen-induced oxygenated compounds vs. herbivore-induced terpenoids) [200,201]. Insect herbivory often increases BVOC emissions [202]; Joutsensaari et al. [203] noted that insect outbreaks could elevate BVOCs and potentially even BSOA emissions. However, severe pest or disease damage may impair physiological functions, such as reduced photosynthesis, resulting in a decrease in the overall production of BVOCs [204]. However, Litvak and Monson [205] found that while monoterpene synthesis in ponderosa pine increased following herbivore induction, the volatile losses from damage issues offset this increase, resulting in no significant net increase in the total monoterpene pool within the damaged needles.

5.2. Environmental Factors

5.2.1. Temperature and Light

Temperature and light are critical environmental factors influencing the emission of BVOCs, as they influence plant enzymatic activity and physiological processes [206]. Generally, increases in temperature and photosynthetically active radiation (PAR) activate the enzymes responsible for BVOCs synthesis [207]. For instance, Li et al. [208] found that high temperatures could triple the monoterpenes emission flux. However, extreme temperatures may disrupt metabolic activities, leading to a decrease in emission rates. Loreto et al. [209] found that the isoprene emission from reed leaves increases with temperatures until 45 °C, after which emissions decline.
Different BVOC components exhibit varying dependencies on temperature and light. The synthesis of isoprene, which primarily occurs in chloroplasts, is directly dependent on the production of glyceraldehyde-3-phosphate (G3P) and energy from photosynthesis. Consequently, isoprene emissions are most influenced by light exposure [210,211,212]. Seasonal and latitudinal variations also affect emissions, with higher light intensities in summer and lower latitudes leading to stronger BVOC emissions.

5.2.2. Humidity

Humidity indirectly influences BVOC emissions by influencing stomata conductance [213,214]. Higher humidity levels promote stomatal opening, facilitating BVOC emissions [215,216]. However, prolonged high humidity can reduce transpiration rates and induce stomatal closure, thereby reducing BVOC emissions [217]. BVOCs’ responses to humidity also vary by components. For instance, Ning et al. [218] found a positive correlation between α-pinene emissions and humidity, while isoprene emissions exhibited a negative correlation.

5.2.3. Concentration of CO2

Changes in carbon dioxide (CO2) concentration influence photosynthesis and metabolic processes, subsequently affecting BVOC emissions. Elevated CO2 concentrations can inhibit BVOC emissions by suppressing photosynthesis, as observed in the reductions in monoterpenes and sesquiterpenes found by Constable et al. [219]. However, certain plants may exhibit increased BVOC synthesis under high CO2 concentrations, as found by Jasoni et al. [220] in onions. The duration of CO2 exposure is also critical. Rapparini et al. [221] found that short-term exposure can significantly suppress the emission of isoprenoids and monoterpenes, whereas long-term exposure had minimal effects.

5.2.4. O3 Stress

The concentration of BVOCs plays a significant role in the formation of O3. Similarly, O3, as a potent oxidant, can influence the release of BVOCs through a series of complex physiological, biochemical, and genetic regulatory mechanisms [222,223,224,225]. Under increased O3 stress, plants may mitigate oxidative damage by increasing the release of BVOCs with strong antioxidant capacities. For example, isoprene emissions increase by 20–50% under acute O3 exposure (100 ppb for 4 h), as demonstrated in Populus tremula fumigation experiments [187]. Llusià et al. [226] also found that plants subjected to O3 fumigation exhibited increased net photosynthetic and total VOC emission rates. Li et al. [227] also discovered that exposure to O3 pollution increased BVOC emissions in the Yellow River Delta and the Yangtze River Delta by 13.9% and 4.9%, respectively. However, plant responses to O3 stress are also influenced by factors such as the exposure duration and the composition of BVOCs. Vo and Faiola [228] found that terpenoid emission rates decreased or remained unchanged after acute O3 exposure, and that sesquiterpenes exhibited a significantly higher response concentration to O3 than monoterpenes. Yuan et al. [75] found that under short-term high-concentration ozone stimulation, the emission of isoprene from leaves that were not severely damaged was promoted. However, prolonged exposure to high concentrations of O3 significantly inhibited isoprene emissions, resulting in a reduction of 40.4%. Additionally, the decrease in isoprene emissions from middle-layer leaves (61.1%) was significantly greater than that from upper-layer leaves (20.9%). Furthermore, this study also found that isoprene emissions were unrelated to stomatal conductance. The interaction between BVOC emissions and O3 is a dynamic process influenced by multiple factors. Understanding the impact of O3 on BVOC emissions requires consideration of multiple intrinsic and environmental variables.

6. Atmospheric Oxidation Mechanisms of BVOCs

The release of BVOCs has a significant impact on atmospheric chemistry. Once released into the atmosphere, these BVOCs undergo a complex series of photochemical reactions that produce and affect the corresponding free radicals, which influence the formation of O3 and SOA in the atmosphere. Understanding the role of BVOCs in atmospheric chemistry and their transformation processes (Figure 5) is critical for comprehending atmospheric chemistry, air quality, and climate change implications [229,230].

6.1. Initial Oxidation Reactions

The initial oxidation reaction pathways of BVOCs in the atmosphere include reactions with hydroxyl radicals (•OH), •NO3, and O3 [231,232,233,234].
The primary source of •OH in the troposphere is from the decomposition of O3 under ultraviolet radiation, which generates excited oxygen atoms O(1D). This process results in high concentrations of •OH during the daytime [235,236,237]. Consequently, during the day when ultraviolet intensity is high, highly reactive BVOCs frequently undergo addition and oxidation reactions with •OH. The hydrogen atoms in carbon-based BVOCs are abstracted by •OH to produce alkyl radicals (R•) and H2O [238,239]. At night, in the absence of light, the reactions between •OH and BVOCs are greatly inhibited, while •NO3, having high stability, becomes the primary radical reacting with BVOCs during nighttime [240,241,242]. Mogensen [243], Xu [244], and Waring and Wells [245] corroborate this mechanism. Since the reaction between O3 and BVOCs is not dependent on light and is highly reactive, at night, •NO3 radicals (derived from O3 + NO2) are typically the primary oxidants for BVOCs in NOx-rich environments. However, in regions with extremely low NOx levels (e.g., pristine forests), the scarcity of •NO3 shifts the dominant oxidation pathway to reactions between O3 and unsaturated BVOCs (e.g., isoprene and monoterpenes), generating carbonyl compounds and sustaining nighttime radical chemistry [246,247,248]. Atkinson [249], Docherty [250], and Kroll [251] found that the oxidation of organic compounds by O3 differs from radical oxidation because BVOCs can be cleaved by O3 to form carbonyls and energetically excited carbonyl oxides. These intermediates then decompose via the hydroperoxide channel to produce •OH and R•. The organic peroxyl radicals (RO2•) from the primary oxidation of BVOCs are part of the radical cycling that controls atmospheric oxidative capacity, leading to the subsequent formation of O3 and SOA [252].

6.2. O3 Formation

Numerous studies have demonstrated that BVOCs play a critical role in the formation of O3 through photochemical reactions, and this role largely depends on the concentration of nitrogen oxides (NOx) [231,253,254,255]. According to the studies by Xie [256] and Kim et al. [257], biogenic and anthropogenic VOCs, especially the photochemical reactions of isoprene, significantly increase the formation of local O3.
Under high-NOx environments, RO2• reacts mainly with NO to produce new oxidation products (RO•, NO2). The newly formed NO2 undergoes photolysis (NO2+ hν → NO + O(3P)), followed by O(3P) + O2 → O3, while RO• decomposes to yield carbonyl compounds (e.g., HCHO) that further sustain radical cycling [258,259]. Simon et al. [260] found that to a certain extent, the more NOx there is, the higher the rate of O3 production, which is largely dependent on the NO2/NO ratio and the amount of BVOCs available. Zhang et al. [261] found that when in high-NOx conditions, the increase in BVOCs will effectively promote radical cycling and enhance the level of radicals, which to some extent may lead to an increase in O3.
In contrast, the reaction of RO2• with hydroperoxy radicals (HO2•) is more important in low NOx concentration environments, where NOx is not a major limiting factor. RO2• reacts with HO2• to produce peroxides (e.g., organic hydroperoxides (ROOH)), and these reactions do not usually result in the direct production of NO2, and therefore contribute less to O3 production. In this environment, the rate of O3 production is limited by the concentration of VOCs and the rate of reaction [262,263]. Orlando and Tyndall [252] and Iyer et al. [264] found that under low NOx conditions, RO2• reacts mainly with HO2•, a reaction that is very important in the atmosphere and is the end of O3 production, manifesting itself in the fact that BVOCs produce less O3 or even consume O3.

6.3. SOA Formation

BVOCs are crucial precursors for the formation of SOA [265,266,267]. SOA consists of fine particulate matter formed through a series of complex chemical transformations of BVOCs, significantly impacting atmospheric climate and human health [268,269]. Among these, isoprene, terpenes, and sesquiterpenes are the most common BVOC precursors [270,271].
The formation process of SOA involves multiple reaction steps. BVOCs first undergo gas-phase oxidation, generating RO2•. These radicals can react through different pathways to produce secondary products, such as reacting with NO to form NO2 and R•, reacting with HO2• to produce ROOH, or reacting with other RO2• to form peroxides (ROOR) and other organics [272,273]. These organics include low-volatility organic compounds (LVOCs) or extremely low-volatility organic compounds (ELVOCs) [274,275]. Concurrently, these LVOCs and ELVOCs can further grow into larger sizes through processes like complexation, isomerization, and condensation nucleation, ultimately forming SOAs [276,277]. This process is also influenced by environmental factors such as temperature, humidity, and NOx concentrations [278,279]. BVOCs are transformed into SOA through complex chemical processes in the atmosphere, influenced by multiple environmental factors. BVOCs such as isoprene, terpenes, and sesquiterpenes undergo oxidation, radical reactions, condensation and adsorption complex reactions, and isomerization, forming SOAs through multiple steps. The formation and evolution of SOA not only affect the balance of atmospheric chemistry but also have significant implications for air quality and human health.

6.4. Synergistic Effects of BVOCs with AVOCs

Both AVOCs (such as NOx and toluene) and BVOCs (such as isoprene and α-pinene) participate in atmospheric photochemical reactions that exacerbate O3 production [280] and promote the condensation of ELVOCs into SOA [281]. In high-NOx environments, the reactivity of BVOCs is enhanced; for instance, BVOCs accounted for 56% of emissions during summer nights in Taipei, thereby extending the photochemical chain when combined with AVOCs. In semi-urban areas of Turkey, the cumulative impact of traffic emissions (AVOCs at 16%) and vegetation releases (BVOCs at 38.6%) has resulted in peak summer O3 levels, with both contributing synergistically to over 60% of ozone formation potential [282,283]. Li et al. [284] found, through analyzing results from different simulation scenarios, that AVOCs promote O3 production by interacting with BVOCs, with the effect being most pronounced from noon to evening. Choi et al. [285] found, through combining Proton Transfer Reaction Time of Flight Mass Spectrometer (PTR-ToF-MS) and the Aerodyne High Resolution Time of Flight Aerosol Mass Spectrometer (HR-ToF-AMS) data, that BVOCs (e.g., monoterpenes) and AVOCs (e.g., toluene) cooperatively drive SOA production through daytime OH oxidation and nighttime NO3 chemistry, with aqueous-phase reactions dominating under high humidity and gas-phase partitioning prevailing in polluted air masses. Ghirardo et al. [286] found that BVOCs (e.g., sesquiterpenes) synergistically enhance SOA formation with AVOCs through NOx-mediated oxidation pathways, where anthropogenic NOx promotes the conversion of both BVOCs and AVOCs into low-volatility products and aerosol aging processes involving multi-phase reactions (e.g., acid-catalyzed polymerization), with their combined contributions amplified under urban pollution and plant stress conditions.
In high-NOx environments, the O3 and SOA formation efficiency of BVOCs is significantly enhanced. Moreover, emission control strategies targeting anthropogenic sources, such as reducing VOCs or NOx, can either amplify or diminish the contribution of biogenic sources. Notably, under carbon neutrality pathways, biogenic contributions may surpass those from anthropogenic sources, highlighting the dual challenges of regional collaborative emission reduction and climate change.

7. Atmospheric Impacts of BVOCs

As critical mediators of biosphere–atmosphere interactions, biogenic volatile organic compounds (BVOCs) profoundly influence air quality, climate systems, and public health through complex atmospheric oxidation pathways [287,288].

7.1. Ozone Formation Potential (OFP) and Climate Feedbacks

The OFP quantifies the mass-specific contribution of BVOCs to ozone production. The O3 formation potentials of isoprene and monoterpenes were 0.05–57.21 and 0.003–72.35 ppb [289]. In addition to this, aromatic hydrocarbons such as toluene and o-xylene and olefins such as ethylene and propylene also contribute significantly to OFP [290]. This potential is amplified in the urban–rural boundary layer, where BVOC-NOx synergies elevate surface O3 by 13–16 ppb [291]. For instance, emissions of reactive olefins (e.g., isoprene) in Zhengzhou during the summer combine with traffic NOx to result in O3 concentrations that exceed China’s Grade II standard (93.3 ppbv) by more than 50% [292].

7.2. Secondary Organic Aerosol Formation Potential (SOAP) and Environmental Impacts

The SOAP characterizes the nucleation efficiency of BVOC oxidation products. The SOA values for isoprene and monoterpenes are 0.0002–0.21 and 0.0006–10.46 μg/m3, respectively [289]. Sesquiterpene oxidation products play an important role in the formation and oxidative chemistry of atmospheric SOA, especially during the spring recovery period [229]. BSOAs exhibit different diurnal formation mechanisms. NOx-induced SOA production dominates at night, and its impact on the environment is influenced by solar radiation, temperature, and humidity [293]. It has been shown that nocturnal terpene ozonolysis with high NOx levels can significantly contribute to secondary organic aerosol formation and organic nitrate formation through anthropogenic–biotic interactions [244].
From 2000 to 2019, the contribution of secondary species to 2.5-micrometer particulate matter (PM2.5) in the Pearl River Delta, China, stabilized at about 80%, with SOAs and secondary inorganic aerosols (SIAs) contributing equally to fine particulate pollution [294]. The presence of SOAs in the atmosphere has implications for climate and health. Respiratory diseases, reduced visibility, haze, deterioration of air quality leading to environmental pollution, etc., are the result of the presence of organic species in atmospheric aerosols [295].

8. Future Research Challenges

While significant progress has been made in understanding BVOC emission mechanisms, characterization methods, and their atmospheric impacts, several challenges remain:
(1) Modeling methods such as Guenther, MEGAN, and BEIS are research methods that overcome spatial and temporal limitations. However, accurate and credible emission prediction models of BVOCs require a substantial amount of high-precision and readily accessible measurement data.
(2) Free radical reactions play a key role in the atmospheric transformation process of BVOCs, directly influencing the levels of O3 and SOA. Nevertheless, the specific reaction pathways in the oxidation stage of BVOCs, as well as the details and mechanisms of the dynamic changes in free radicals, remain poorly understood.
(3) The emission of BVOCs is affected by a multitude of factors, many of which do not individually influence the emission of BVOCs. There is a dearth of systematic studies on the multiple synergistic effects of the internal and external factors on the emission mechanisms of BVOCs and their long-term impacts.
Given these challenges, the following three research areas are particularly important for future BVOC studies: (1) the development of high-precision, low-cost, and easy-to-operate methods for rapid measurements of BVOCs; (2) an in-depth study of the atmospheric chemical reaction mechanism between BVOCs and free radicals; and (3) an investigation into the synergistic effects of multiple factors on the emission mechanism of BVOCs.

9. Recommendations and Prospects

9.1. Recommendations of BVOC Prevention and Control

(1) Formulate BVOC management strategies or guidelines:
Among the world’s major economies, the EU’s Clean Air Quality Directive, the U.S. Clean Air Act, and China’s Comprehensive Control Program for Volatile Organic Compounds in Key Industries all have stringent controls on the emission of anthropogenic VOCs. However, all of these programs merge and unify the management of anthropogenic sources of VOCs, such as vehicles, industry, and construction materials, and do not provide effective guidance on BVOCs. Meanwhile, the guidance of each country is basically formulated based on its own situation and lacks universality and popularization. It is recommended that developed countries with high forest cover, such as the U.S. and Canada, and developing countries with high forest resources, such as China and Brazil, cooperate to incorporate the management of BVOCs into an integrated air quality management framework, focusing on their synergistic effects with anthropogenic pollutants, and to further improve the global VOCs pollution situation.
(2) Establish BVOC emission control indicators based on multi-dimensional parameters:
In contrast to the relatively straightforward and predictable emissions observed in laboratory settings, the release of BVOCs in natural environments is a complex and variable process. It is therefore evident that a single indicator cannot be relied upon for the quantitative control of BVOC emissions. By cross-analyzing multi-dimensional parameters through big data, accurately identifying the emission characteristics of different BVOC sources, and highlighting the major items, a more scientific and targeted BVOC emission control indicator can be formulated. This approach helps to avoid the inaccuracy and significant errors of single indicators, optimize the use of resources, and reduce unnecessary economic losses and control costs.
(3) Dynamically adjust BVOC management based on temporal and spatial characteristics of BVOC emissions:
BVOC emissions exhibit significant seasonality and diurnal variations over time and are influenced by geographical and climatic conditions spatially. Additionally, the synergistic pollution effects with other pollutants like anthropogenic NOx vary. Therefore, it is essential to implement differentiated control strategies based on the temporal and spatial characteristics of specific regions and adjust BVOC management strategies dynamically according to local conditions. This makes policies more flexible and scientific, enhancing the effectiveness and specificity of regional management.

9.2. Prospects of BVOC Monitor Strategies

(1) Develop more efficient, convenient, and low-cost BVOC measurement technologies:
As previously stated, a rapid sampling method for BVOCs that is highly accurate, straightforward to operate, and cost-effective is the foundation for obtaining high-quality detection data and conducting related research. The Guangzhou Institute of Geochemistry, Chinese Academy of Sciences (GIGC) devised a semi-open dynamic chamber system for measuring BVOCs. This system can reduce the time required to reach a steady state at higher flow rates, thereby minimizing adsorption loss. Nevertheless, the adsorption loss for macromolecular compounds remains above 30%, indicating a need for further optimization [296]. Electrochemical sensors are employed for the measurement of atmospheric NO2 due to their low cost, portability, and flexibility, which can also be applied to the measurement of BVOCs [297]. Laser-induced fluorescence tracking has also been employed for the sampling and measurement of BVOCs, although the technology is not yet fully developed [298]. It is therefore feasible to pursue the optimization of existing methods or the adoption of techniques from other fields as avenues for the development of more efficient, convenient, and cost-effective measurement techniques for BVOCs.
(2) Develop BVOC monitoring and prediction technologies based on high temporal and spatial resolution big data models:
The advent of new information science and technology has opened up new avenues for research in the field of environmental management. By extending the aforementioned laser-induced fluorescence tracking technology, the regional dissemination of environmentally friendly fluorescent probes, and the laser-induced generation of remotely sensed data through the use of UAVs, in conjunction with ground-based observation data, a comprehensive database of high temporal and spatial resolution BVOC emissions can be formed. The application of machine learning and deep learning algorithms can enhance the synergistic effects of multiple factors, optimize existing BVOC emission models, and facilitate the development of new prediction models. A map of BVOC emission characteristics in different regions can be constructed, allowing region-specific control strategies to be formulated and precise management to be achieved.
(3) Promote the selection of low-BVOC-emitting tree species in urban greening and ecological restoration:
In urban greening and ecological restoration, besides considering the ornamental value and ecological function of trees, selecting low-BVOC-emitting tree species can help optimize air quality while maintaining ecological function and aesthetic values.
(4) Artificially control precursor substance emissions:
The strengthening of emission controls of precursors such as O3 and SOA can indirectly reduce the emission level of BVOCs, reducing secondary chemical reactions that lead to O3 and SOA formation.

Supplementary Materials

The following Supporting Information can be downloaded at: https://www.mdpi.com/article/10.3390/toxics13050364/s1, Table S1: Nomenclature.

Author Contributions

Conceptualization, J.L.; methodology, R.L., R.G. and Y.Y.; software, Y.Y.; validation, L.W. and M.H.-A.-M.; formal analysis, X.X.; investigation, R.L., L.W. and X.S.; resources, R.G. and H.L.; data curation, R.L., X.S. and X.X.; writing—original draft preparation, R.L.; writing—review and editing, X.L. and R.G.; visualization, M.H.-A.-M.; supervision, X.L. and H.L.; project administration, J.L.; funding acquisition, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (Grant No. 42307149) “Studies on gradient observation of O3 and its precursors based on plateau background mountain stations”, awarded to Dr. Yuan Yang (Principal Investigator).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Acknowledgments

This work is supported by the National Natural Science Foundation of China (Grant No. 42307149).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Cai, M.; An, C.; Guy, C. A scientometric analysis and review of biogenic volatile organic compound emissions: Research hotspots, new frontiers, and environmental implications. Renew. Sustain. Energy Rev. 2021, 149, 111317. [Google Scholar] [CrossRef]
  2. Guenther, A.; Karl, T.; Harley, P.; Wiedinmyer, C.; Palmer, P.I.; Geron, C. Estimates of global terrestrial isoprene emissions using MEGAN (Model of Emissions of Gases and Aerosols from Nature). Atmos. Chem. Phys. 2006, 6, 3181–3210. [Google Scholar] [CrossRef]
  3. Guenther, A.; Hewitt, C.N.; Erickson, D.; Fall, R.; Geron, C.; Graedel, T.; Harley, P.; Klinger, L.; Lerdau, M.; Mckay, W.A.; et al. A global model of natural volatile organic compound emissions. J. Geophys. Res. 1995, 100, 8873. [Google Scholar] [CrossRef]
  4. World Health Organization. WHO Global Air Quality Guidelines: Particulate Matter (PM2.5 and PM10), Ozone, Nitrogen Dioxide, Sulfur Dioxide and Carbon Monoxide; World Health Organization: Geneva, Switzerland, 2021; Available online: https://www.who.int/publications/i/item/9789240034228 (accessed on 27 April 2025).
  5. Eastman, J.R.; Sangermano, F.; Machado, E.A.; Rogan, J.; Anyamba, A. Global Trends in Seasonality of Normalized Difference Vegetation Index (NDVI), 1982–2011. Remote Sens. 2013, 5, 4799–4818. [Google Scholar] [CrossRef]
  6. Huang, L.; Zhao, X.; Chen, C.; Tan, J.; Li, Y.; Chen, H.; Wang, Y.; Li, L.; Guenther, A.; Huang, H. Uncertainties of biogenic VOC emissions caused by land cover data and implications on ozone mitigation strategies for the Yangtze river Delta region. Atmos. Environ. 2024, 337, 120765. [Google Scholar] [CrossRef]
  7. Steinbrecher, R.; Smiatek, G.; Köble, R.; Seufert, G.; Theloke, J.; Hauff, K.; Ciccioli, P.; Vautard, R.; Curci, G. Intra- and inter-annual variability of VOC emissions from natural and semi-natural vegetation in Europe and neighbouring countries. Atmos. Environ. 2009, 43, 1380–1391. [Google Scholar] [CrossRef]
  8. Wang, P.; Chen, Y.; Hu, J.; Zhang, H.; Ying, Q. Source apportionment of summertime ozone in China using a source-oriented chemical transport model. Atmos. Environ. 2019, 211, 79–90. [Google Scholar] [CrossRef]
  9. Gu, Y.; Liu, B.; Meng, H.; Song, S.; Dai, Q.; Shi, L.; Feng, Y.; Hopke, P.K. Source apportionment of consumed volatile organic compounds in the atmosphere. J. Hazard. Mater. 2023, 459, 132138. [Google Scholar] [CrossRef]
  10. Hallquist, M.; Wenger, J.C.; Baltensperger, U.; Rudich, Y.; Simpson, D.; Claeys, M.; Dommen, J.; Donahue, N.M.; George, C.; Goldstein, A.H.; et al. The formation, properties and impact of secondary organic aerosol: Current and emerging issues. Atmos. Chem. Phys. 2009, 9, 5155–5236. [Google Scholar] [CrossRef]
  11. Cao, J.; Situ, S.; Hao, Y.; Xie, S.; Li, L. Enhanced summertime ozone and SOA from biogenic volatile organic compound (BVOC) emissions due to vegetation biomass variability during 1981–2018 in China. Atmos. Chem. Phys. 2022, 22, 2351–2364. [Google Scholar] [CrossRef]
  12. Qin, M.; Hu, Y.; Wang, X.; Vasilakos, P.; Boyd, C.M.; Xu, L.; Song, Y.; Ng, N.L.; Nenes, A.; Russell, A.G. Modeling biogenic secondary organic aerosol (BSOA) formation from monoterpene reactions with NO3: A case study of the SOAS campaign using CMAQ. Atmos. Environ. 2018, 184, 146–155. [Google Scholar] [CrossRef]
  13. Rap, A.; Scott, C.E.; Reddington, C.L.; Mercado, L.; Ellis, R.J.; Garraway, S.; Evans, M.J.; Beerling, D.J.; MacKenzie, A.R.; Hewitt, C.N.; et al. Enhanced global primary production by biogenic aerosol via diffuse radiation fertilization. Nature Geosci. 2018, 11, 640–644. [Google Scholar] [CrossRef]
  14. Yang, K.; Llusià, J.; Preece, C.; Ogaya, R.; Márquez Tur, L.; Mu, Z.; You, C.; Xu, Z.; Tan, Y.; Peñuelas, J. Impacts of seasonality, drought, nitrogen fertilization, and litter on soil fluxes of biogenic volatile organic compounds in a Mediterranean forest. Sci. Total Environ. 2024, 906, 167354. [Google Scholar] [CrossRef] [PubMed]
  15. Sanadze, G.A. Emission of Gaseous Organic Sub-stances from Plants. Rep. Akad. Nauk. GruzSSR 1956, 17, 429–433. [Google Scholar]
  16. Sanadze, G.A. Biogenic Isoprene (A Review). Russ. J. Plant Physiol. 2004, 51, 729–741. [Google Scholar] [CrossRef]
  17. Laothawornkitkul, J.; Taylor, J.E.; Paul, N.D.; Hewitt, C.N. Biogenic volatile organic compounds in the Earth system. New Phytol. 2009, 183, 27–51. [Google Scholar] [CrossRef]
  18. Peng, Q.; Yang, Y.; Ou, W.; Wei, L.; Li, Z.; Deng, X.; Gao, Q. The characteristics and environmental significance of BVOCs released by aquatic macrophytes. Chemosphere 2024, 361, 142574. [Google Scholar] [CrossRef]
  19. Peñuelas, J.; Llusià, J. BVOCs: Plant defense against climate warming? Trends Plant Sci. 2003, 8, 105–109. [Google Scholar] [CrossRef]
  20. Kesselmeier, J.; Staudt, M. Biogenic Volatile Organic Compounds (VOC): An Overview on Emission, Physiology and Ecology. J. Atmos. Chem. 1999, 33, 23–88. [Google Scholar] [CrossRef]
  21. Helmig, D.; Klinger, L.F.; Guenther, A.; Vierling, L.; Geron, C.; Zimmerman, P. Biogenic volatile organic compound emissions (BVOCs) II. Landscape flux potentials from three continental sites in the U.S. Chemosphere 1999, 38, 2189–2204. [Google Scholar] [CrossRef]
  22. Otter, L.B.; Guenther, A.; Greenberg, J. Seasonal and spatial variations in biogenic hydrocarbon emissions from southern African savannas and woodlands. Atmos. Environ. 2002, 36, 4265–4275. [Google Scholar] [CrossRef]
  23. Papiez, M.R.; Potosnak, M.J.; Goliff, W.S.; Guenther, A.B.; Matsunaga, S.N.; Stockwell, W.R. The impacts of reactive terpene emissions from plants on air quality in Las Vegas, Nevada. Atmos. Environ. 2009, 43, 4109–4123. [Google Scholar] [CrossRef]
  24. Kramshøj, M.; Vedel-Petersen, I.; Schollert, M.; Rinnan, Å.; Nymand, J.; Ro-Poulsen, H.; Rinnan, R. Large increases in Arctic biogenic volatile emissions are a direct effect of warming. Nature Geosci. 2016, 9, 349–352. [Google Scholar] [CrossRef]
  25. Guenther, A.B.; Zimmerman, P.R.; Harley, P.C.; Monson, R.K.; Fall, R. Isoprene and monoterpene emission rate variability: Model evaluations and sensitivity analyses. J. Geophys. Res. Atmos. 1993, 98, 12609–12617. [Google Scholar] [CrossRef]
  26. Guenther, A.B.; Monson, R.K.; Fall, R. Isoprene and monoterpene emission rate variability: Observations with eucalyptus and emission rate algorithm development. J. Geophys. Res. Atmos. 1991, 96, 10799–10808. [Google Scholar] [CrossRef]
  27. Pierce, T.E. PC-BEIS: A personal computer version of the biogenic emissions inventory system. J. Air Waste Manag. Assoc. 1991, 41, 937–941. [Google Scholar] [CrossRef]
  28. Ng, N.L.; Brown, S.S.; Archibald, A.T.; Atlas, E.; Cohen, R.C.; Crowley, J.N.; Day, D.A.; Donahue, N.M.; Fry, J.L.; Fuchs, H.; et al. Nitrate radicals and biogenic volatile organic compounds: Oxidation, mechanisms, and organic aerosol. Atmos. Chem. Phys. 2017, 17, 2103–2162. [Google Scholar] [CrossRef]
  29. Tang, J.; Zhou, P.; Miller, P.A.; Schurgers, G.; Gustafson, A.; Makkonen, R.; Fu, Y.H.; Rinnan, R. High-latitude vegetation changes will determine future plant volatile impacts on atmospheric organic aerosols. npj Clim. Atmos. Sci. 2023, 6, 1–13. [Google Scholar] [CrossRef]
  30. Lun, X.; Lin, Y.; Chai, F.; Fan, C.; Li, H.; Liu, J. Reviews of emission of biogenic volatile organic compounds (BVOCs) in Asia. J. Environ. Sci. 2020, 95, 266–277. [Google Scholar] [CrossRef]
  31. Yang, W.; Cao, J.; Wu, Y.; Kong, F.; Li, L. Review on plant terpenoid emissions worldwide and in China. Sci. Total Environ. 2021, 787, 147454. [Google Scholar] [CrossRef]
  32. Smith, D.; Španěl, P. Direct, rapid quantitative analyses of BVOCs using SIFT-MS and PTR-MS obviating sample collection. TrAC Trends Anal. Chem. 2011, 30, 945–959. [Google Scholar] [CrossRef]
  33. Bacsik, Z.; Mink, J.; Keresztury, G. FTIR Spectroscopy of the Atmosphere. I. Principles and Methods. Appl. Spectrosc. Rev. 2004, 39, 295–363. [Google Scholar]
  34. Gómez, M.C.; Durana, N.; García, J.A.; de Blas, M.; Sáez de Cámara, E.; García-Ruiz, E.; Gangoiti, G.; Torre-Pascual, E.; Iza, J. Long-term measurement of biogenic volatile organic compounds in a rural background area: Contribution to ozone formation. Atmos. Environ. 2020, 224, 117315. [Google Scholar]
  35. Kalalian, C.; Roth, E.; El Dib, G.; Singh, H.J.; Rao, P.K.; Chakir, A. Product investigation of the gas phase ozonolysis of 1-penten-3-ol, cis-2-penten-1-ol and trans-3-hexen-1-ol. Atmos. Environ. 2020, 238, 117732. [Google Scholar] [CrossRef]
  36. Shi, Y.; Wang, B.; Yin, C.; Li, Z.; Yu, Y. Performance improvement: A lightweight gas information classification method combined with an electronic nose system. Sens. Actuators B Chem. 2023, 396, 134551. [Google Scholar] [CrossRef]
  37. Das, P.P. Glass microfiber based method for extraction and detection of BVOCs of Camellia sinensis (L.) O. Kuntze. J. Appl. Res. Med. Aromat. Plants 2021, 24, 100319. [Google Scholar]
  38. Fattobene, M.; Papa, F.; Russo, R.E.; Zamponi, S.; Conti, P.; Taffetani, F.; Sorci, A.; Liu, F.; Berrettoni, M. ON-SITE monitoring OF BVOCS emission in Tremiti island, Italy. Heliyon 2024, 10, e23822. [Google Scholar] [CrossRef]
  39. Baroja, O.; Unceta, N.; Sampedro, M.C.; Goicolea, M.A.; Barrio, R.J. Optimization and validation of a method of analysis for fenitrothion and its main metabolites in forestry air samples using sorbent tubes with thermal desorption cold trap injection and gas chromatography–mass spectrometry. J. Chromatogr. A 2004, 1059, 165–170. [Google Scholar] [CrossRef]
  40. Wu, K.; Yang, X.; Chen, D.; Gu, S.; Lu, Y.; Jiang, Q.; Wang, K.; Ou, Y.; Qian, Y.; Shao, P.; et al. Estimation of biogenic VOC emissions and their corresponding impact on ozone and secondary organic aerosol formation in China. Atmos. Res. 2020, 231, 104656. [Google Scholar] [CrossRef]
  41. Gonçalves, O.C.; Cerqueira, J.S.R.F.; Mestre, A.S.; Neng, N.R.; Nogueira, J.M.F. HS-BAμE: A New Alternative Approach for VOCs Analysis—Application for Monitoring Biogenic Emissions from Tree Species. Molecules 2023, 28, 1179. [Google Scholar] [CrossRef]
  42. Kim, S.; Karl, T.; Guenther, A.; Tyndall, G.; Orlando, J.; Harley, P.; Rasmussen, R.; Apel, E. Emissions and ambient distributions of Biogenic Volatile Organic Compounds (BVOC) in a ponderosa pine ecosystem: Interpretation of PTR-MS mass spectra. Atmos. Chem. Phys. 2010, 10, 1759–1771. [Google Scholar] [CrossRef]
  43. Lindinger, W.; Jordan, A. Proton-transfer-reaction mass spectrometry (PTR–MS): On-line monitoring of volatile organic compounds at pptv levels. Chem. Soc. Rev. 1998, 27, 347–375. [Google Scholar] [CrossRef]
  44. D’Arco, A.; Mancini, T.; Paolozzi, M.C.; Macis, S.; Mosesso, L.; Marcelli, A.; Petrarca, M.; Radica, F.; Tranfo, G.; Lupi, S.; et al. High Sensitivity Monitoring of VOCs in Air through FTIR Spectroscopy Using a Multipass Gas Cell Setup. Sensors 2022, 22, 5624. [Google Scholar] [CrossRef] [PubMed]
  45. Fathy, A.; Sabry, Y.M.; Amr, M.; Gnambodoe-Capo-chichi, M.; Anwar, M.; Ghoname, A.O.; Amr, A.; Saeed, A.; Gad, M.; Haron, M.A.; et al. MEMS FTIR optical spectrometer enables detection of volatile organic compounds (VOCs) in part-per-billion (ppb) range for air quality monitoring. In MOEMS and Miniaturized Systems XVIII, Proceedings of the SPIE OPTO, San Francisco, CA, USA, 2–7 February 2019; SPIE: Bellingham, WA, USA; Volume 10931, pp. 69–75.
  46. Mermet, K.; Sauvage, S.; Dusanter, S.; Salameh, T.; Léonardis, T.; Flaud, P.-M.; Perraudin, É.; Villenave, E.; Locoge, N. Optimization of a gas chromatographic unit for measuring biogenic volatile organic compounds in ambient air. Atmos. Meas. Tech. 2019, 12, 6153–6171. [Google Scholar] [CrossRef]
  47. Chang, Y.; Tang, N.; Qu, H.; Liu, J.; Zhang, D.; Zhang, H.; Pang, W.; Duan, X. Detection of Volatile Organic Compounds by Self-assembled Monolayer Coated Sensor Array with Concentration-independent Fingerprints. Sci. Rep. 2016, 6, 23970. [Google Scholar] [CrossRef]
  48. Fair, J.D.; Bailey, W.F.; Felty, R.A.; Gifford, A.E.; Shultes, B.; Volles, L.H. Method for rapid on-site identification of VOCs. J. Environ. Sci. 2009, 21, 1005–1008. [Google Scholar] [CrossRef]
  49. Stierlin, É.; Michel, T.; Fernandez, X. Field analyses of lavender volatile organic compounds: Performance evaluation of a portable gas chromatography–mass spectrometry device. Phytochem. Anal. 2020, 31, 778–785. [Google Scholar] [CrossRef]
  50. Santos, F.J.; Galceran, M.T. The application of gas chromatography to environmental analysis. TrAC Trends Anal. Chem. 2002, 21, 672–685. [Google Scholar] [CrossRef]
  51. Santhanakrishnan, V.P.; Shoba, N.; Senthamizh Selvi, B.; Varun, E.; Mohankumar, S.; Raveendran, M. Aromatic profiling of Murraya koenigii leaves by Thermal Desorption Gas chromatography-Mass Spectroscopy (TD-GC-MS). Heliyon 2023, 9, e17832. [Google Scholar]
  52. Ren, Q.; Jin, Y.; Hu, Y.; Chen, H.; Li, Z. Rapid changes of induced volatile organic compounds in Pinus massoniana. Front. Forest. China 2007, 2, 459–465. [Google Scholar] [CrossRef]
  53. Kim, Y.-H.; Kim, K.-H. Ultimate Detectability of Volatile Organic Compounds: How Much Further Can We Reduce Their Ambient Air Sample Volumes for Analysis? Anal. Chem. 2012, 84, 8284–8293. [Google Scholar] [CrossRef] [PubMed]
  54. Lee, J.-H.; Hwang, S.-M.; Lee, D.-W.; Heo, G.-S. Determination of Volatile Organic Compounds (VOCs) Using Tedlar Bag/Solid-phase Microextraction/Gas Chromatography/Mass Spectrometry (SPME/GC/MS) in Ambient and Workplace Air. Bull. Korean Chem. Soc. 2002, 23, 488–496. [Google Scholar]
  55. Feng, J.; Zhang, B.; Zhang, H.; Wu, Z.; Li, M.; Wang, D.; Wang, C. Combining with E-nose, GC-MS, GC-IMS and chemometrics to explore volatile characteristics during the different stages of Zanthoxylum bungeanum maxim fruits. Food Res. Int. 2024, 195, 114964. [Google Scholar] [CrossRef]
  56. Wang, W.; Liu, C.; Clarisse, L.; Van Damme, M.; Coheur, P.-F.; Xie, Y.; Shan, C.; Hu, Q.; Sun, Y.; Jones, N. Ground-based measurements of atmospheric NH3 by Fourier transform infrared spectrometry at Hefei and comparisons with IASI data. Atmos. Environ. 2022, 287, 119256. [Google Scholar] [CrossRef]
  57. Yuan, B.; Koss, A.R.; Warneke, C.; Coggon, M.; Sekimoto, K.; de Gouw, J.A. Proton-Transfer-Reaction Mass Spectrometry: Applications in Atmospheric Sciences. Chem. Rev. 2017, 117, 13187–13229. [Google Scholar] [CrossRef]
  58. Hansel, A.; Jordan, A.; Holzinger, R.; Prazeller, P.; Vogel, W.; Lindinger, W. Proton transfer reaction mass spectrometry: On-line trace gas analysis at the ppb level. Int. J. Mass Spectrom. Ion Process. 1995, 149–150, 609–619. [Google Scholar] [CrossRef]
  59. Zhu, B.; Han, Y.; Wang, C.; Huang, X.; Xia, S.; Niu, Y.; Yin, Z.; He, L. Understanding primary and secondary sources of ambient oxygenated volatile organic compounds in Shenzhen utilizing photochemical age-based parameterization method. J. Environ. Sci. 2019, 75, 105–114. [Google Scholar] [CrossRef] [PubMed]
  60. Bessala, L.F.B.; Gao, J.; He, Z.; Wang, Z.; Yi, S. Effects of different heat treatment media on odorous constituents, chemical decomposition and mechanical properties of two hardwoods. RSC Adv. 2024, 14, 7414–7429. [Google Scholar] [CrossRef]
  61. Fathy, A.; Pivert, M.L.; Kim, Y.J.; Ba, M.O.; Erfan, M.; Sabry, Y.M.; Khalil, D.; Leprince-Wang, Y.; Bourouina, T.; Gnambodoe-Capochichi, M. Continuous Monitoring of Air Purification: A Study on Volatile Organic Compounds in a Gas Cell. Sensors 2020, 20, 934. [Google Scholar] [CrossRef]
  62. Ma, F.; Zhang, G.; Guo, H.; Liao, L.; Huang, X.; Yi, Z. Transient interaction effects of temperature and light intensity on isoprene and monoterpene emissions from Schima superba and Phoebe bournei. Sci. Total Environ. 2023, 894, 165082. [Google Scholar] [CrossRef]
  63. Rhew, R.C.; Deventer, M.J.; Turnipseed, A.A.; Warneke, C.; Ortega, J.; Shen, S.; Martinez, L.; Koss, A.; Lerner, B.M.; Gilman, J.B.; et al. Ethene, propene, butene and isoprene emissions from a ponderosa pine forest measured by relaxed eddy accumulation. Atmos. Chem. Phys. 2017, 17, 13417–13438. [Google Scholar] [CrossRef]
  64. Copolovici, L.; Niinemets, Ü. Flooding induced emissions of volatile signalling compounds in three tree species with differing waterlogging tolerance. Plant Cell Environ. 2010, 33, 1582–1594. [Google Scholar] [CrossRef]
  65. You, D.-W.; Seon, Y.-S.; Jang, Y.; Bang, J.; Oh, J.-S.; Jung, K.-W. A portable gas chromatograph for real-time monitoring of aromatic volatile organic compounds in air samples. J. Chromatogr. A 2020, 1625, 461267. [Google Scholar] [CrossRef] [PubMed]
  66. Sharma, R.; Zhou, M.; Hunter, M.D.; Fan, X. Rapid In Situ Analysis of Plant Emission for Disease Diagnosis Using a Portable Gas Chromatography Device. J. Agric. Food Chem. 2019, 67, 7530–7537. [Google Scholar] [CrossRef] [PubMed]
  67. Ciccioli, P.; Silibello, C.; Finardi, S.; Pepe, N.; Ciccioli, P.; Rapparini, F.; Neri, L.; Fares, S.; Brilli, F.; Mircea, M.; et al. The potential impact of biogenic volatile organic compounds (BVOCs) from terrestrial vegetation on a Mediterranean area using two different emission models. Agric. For. Meteorol. 2023, 328, 109255. [Google Scholar] [CrossRef]
  68. Anisimov, D.S.; Chekusova, V.P.; Trul, A.A.; Abramov, A.A.; Borshchev, O.V.; Agina, E.V.; Ponomarenko, S.A. Fully integrated ultra-sensitive electronic nose based on organic field-effect transistors. Sci. Rep. 2021, 11, 10683. [Google Scholar] [CrossRef]
  69. Feng, X.; Wang, H.; Wang, Z.; Huang, P.; Kan, J. Discrimination and characterization of the volatile organic compounds in eight kinds of huajiao with geographical indication of China using electronic nose, HS-GC-IMS and HS-SPME-GC–MS. Food Chem. 2022, 375, 131671. [Google Scholar] [CrossRef]
  70. Lin, C.-Y.; Chang, T.-C.; Chen, Y.-H.; Chen, Y.-J.; Cheng, S.-S.; Chang, S.-T. Monitoring the dynamic emission of biogenic volatile organic compounds from Cryptomeria japonica by enclosure measurement. Atmos. Environ. 2015, 122, 163–170. [Google Scholar] [CrossRef]
  71. Li, L.; Guenther, A.B.; Xie, S.; Gu, D.; Seco, R.; Nagalingam, S.; Yan, D. Evaluation of semi-static enclosure technique for rapid surveys of biogenic volatile organic compounds (BVOCs) emission measurements. Atmos. Environ. 2019, 212, 1–5. [Google Scholar] [CrossRef]
  72. Aydin, Y.M.; Yaman, B.; Koca, H.; Dasdemir, O.; Kara, M.; Altiok, H.; Dumanoglu, Y.; Bayram, A.; Tolunay, D.; Odabasi, M.; et al. Biogenic volatile organic compound (BVOC) emissions from forested areas in Turkey: Determination of specific emission rates for thirty-one tree species. Sci. Total Environ. 2014, 490, 239–253. [Google Scholar] [CrossRef]
  73. Vedel-Petersen, I.; Schollert, M.; Nymand, J.; Rinnan, R. Volatile organic compound emission profiles of four common arctic plants. Atmos. Environ. 2015, 120, 117–126. [Google Scholar] [CrossRef]
  74. Jing, X.; Lun, X.; Fan, C.; Ma, W. Emission patterns of biogenic volatile organic compounds from dominant forest species in Beijing, China. J. Environ. Sci. 2020, 95, 73–81. [Google Scholar] [CrossRef]
  75. Yuan, X.; Calatayud, V.; Gao, F.; Fares, S.; Paoletti, E.; Tian, Y.; Feng, Z. Interaction of drought and ozone exposure on isoprene emission from extensively cultivated poplar. Plant Cell Environ. 2016, 39, 2276–2287. [Google Scholar] [CrossRef]
  76. Beauchamp, J.; Herbig, J.; Gutmann, R.; Hansel, A. On the use of Tedlar® bags for breath-gas sampling and analysis. J. Breath Res. 2008, 2, 046001. [Google Scholar] [CrossRef] [PubMed]
  77. Hsieh, C.-C.; Horng, S.-H.; Liao, P.-N. Stability of Trace-level Volatile Organic Compounds Stored in Canisters and Tedlar Bags. Aerosol Air Qual. Res. 2003, 3, 17–28. [Google Scholar] [CrossRef]
  78. Arnts, R.R. Evaluation of adsorbent sampling tube materials and Tenax-TA for analysis of volatile biogenic organic compounds. Atmos. Environ. 2010, 44, 1579–1584. [Google Scholar] [CrossRef]
  79. Hellén, H.; Tykkä, T.; Schallhart, S.; Stratigou, E.; Salameh, T.; Iturrate-Garcia, M. Measurements of atmospheric C10–C15 biogenic volatile organic compounds (BVOCs) with sorbent tubes. Atmos. Meas. Tech. 2024, 17, 315–333. [Google Scholar] [CrossRef]
  80. Nassiri, P.; Golbabaie, F.; Mehrain, K.; Hematian, A. Evaluation of activated charcoal sampler tubes. Iran. J. Public Health 1994, 23, 11–20. [Google Scholar]
  81. Prado, C.; Alcaraz, M.J.; Fuentes, A.; Garrido, J.; Periago, J.F. Storage stability of ketones on carbon adsorbents. J. Chromatogr. A 2006, 1129, 82–87. [Google Scholar] [CrossRef]
  82. Wang, D.K.W.; Austin, C.C. Determination of complex mixtures of volatile organic compounds in ambient air: Canister methodology. Anal. Bioanal. Chem. 2006, 386, 1099–1120. [Google Scholar] [CrossRef]
  83. Acton, W.J.F.; Schallhart, S.; Langford, B.; Valach, A.; Rantala, P.; Fares, S.; Carriero, G.; Tillmann, R.; Tomlinson, S.J.; Dragosits, U.; et al. Canopy-scale flux measurements and bottom-up emission estimates of volatile organic compounds from a mixed oak and hornbeam forest in northern Italy. Atmos. Chem. Phys. 2016, 16, 7149–7170. [Google Scholar] [CrossRef]
  84. Karl, T.; Misztal, P.K.; Jonsson, H.H.; Shertz, S.; Goldstein, A.H.; Guenther, A.B. Airborne Flux Measurements of BVOCs above Californian Oak Forests: Experimental Investigation of Surface and Entrainment Fluxes, OH Densities, and Damkohler Numbers. J. Atmos. Sci. 2013, 70, 3277–3287. [Google Scholar] [CrossRef]
  85. Baldocchi, D.D.; Hincks, B.B.; Meyers, T.P. Measuring Biosphere-Atmosphere Exchanges of Biologically Related Gases with Micrometeorological Methods. Ecology 1988, 69, 1331–1340. [Google Scholar] [CrossRef]
  86. Spirig, C.; Neftel, A.; Ammann, C.; Dommen, J.; Grabmer, W.; Thielmann, A.; Schaub, A.; Beauchamp, J.; Wisthaler, A.; Hansel, A. Eddy covariance flux measurements of biogenic VOCs during ECHO 2003 using proton transfer reaction mass spectrometry. Atmos. Chem. Phys. 2005, 5, 465–481. [Google Scholar] [CrossRef]
  87. Misztal, P.K.; Nemitz, E.; Langford, B.; Di Marco, C.F.; Phillips, G.J.; Hewitt, C.N.; MacKenzie, A.R.; Owen, S.M.; Fowler, D.; Heal, M.R.; et al. Direct ecosystem fluxes of volatile organic compounds from oil palms in South-East Asia. Atmos. Chem. Phys. 2011, 11, 8995–9017. [Google Scholar] [CrossRef]
  88. Bai, J.; Guenther, A.; Turnipseed, A.; Duhl, T.; Greenberg, J. Seasonal and interannual variations in whole-ecosystem BVOC emissions from a subtropical plantation in China. Atmos. Environ. 2017, 161, 176–190. [Google Scholar] [CrossRef]
  89. Park, J.-H.; Fares, S.; Weber, R.; Goldstein, A.H. Biogenic volatile organic compound emissions during BEARPEX 2009 measured by eddy covariance and flux-gradient similarity methods. Atmos. Chem. Phys. 2014, 14, 231–244. [Google Scholar] [CrossRef]
  90. Holst, T.; Arneth, A.; Hayward, S.; Ekberg, A.; Mastepanov, M.; Jackowicz-Korczynski, M.; Friborg, T.; Crill, P.M.; Backstrand, K. BVOC ecosystem flux measurements at a high latitude wetland site. Atmos. Chem. Phys. 2010, 10, 1617–1634. [Google Scholar] [CrossRef]
  91. Bachy, A.; Aubinet, M.; Schoon, N.; Amelynck, C.; Bodson, B.; Moureaux, C.; Heinesch, B. Are BVOC exchanges in agricultural ecosystems overestimated? Insights from fluxes measured in a maize field over a whole growing season. Atmos. Chem. Phys. 2016, 16, 5343–5356. [Google Scholar] [CrossRef]
  92. Edwards, G.D.; Martins, D.K.; Starn, T.; Pratt, K.; Shepson, P.B. A disjunct eddy accumulation system for the measurement of BVOC fluxes: Instrument characterizations and field deployment. Atmos. Meas. Tech. 2012, 5, 2115–2132. [Google Scholar] [CrossRef]
  93. Chen, Y.; Zhang, X.; Wu, X.; Li, J.; Qiu, Y.; Wang, H.; Cheng, Z.; Zheng, C.; Yang, F. Volatile Organic Compound Sampling through Rotor Unmanned Aerial Vehicle Technique for Environmental Monitoring. Atmosphere 2022, 13, 1442. [Google Scholar] [CrossRef]
  94. Li, Y.; Liu, B.; Ye, J.; Jia, T.; Khuzestani, R.B.; Sun, J.Y.; Cheng, X.; Zheng, Y.; Li, X.; Wu, C.; et al. Unmanned Aerial Vehicle Measurements of Volatile Organic Compounds over a Subtropical Forest in China and Implications for Emission Heterogeneity. ACS Earth Space Chem. 2021, 5, 247–256. [Google Scholar] [CrossRef]
  95. McKinney, K.A.; Wang, D.; Ye, J.; de Fouchier, J.-B.; Guimarães, P.C.; Batista, C.E.; Souza, R.A.F.; Alves, E.G.; Gu, D.; Guenther, A.B.; et al. A sampler for atmospheric volatile organic compounds by copter unmanned aerial vehicles. Atmos. Meas. Tech. 2019, 12, 3123–3135. [Google Scholar] [CrossRef]
  96. Ashworth, K.; Boissard, C.; Folberth, G.; Lathière, J.; Schurgers, G. Global Modelling of Volatile Organic Compound Emissions. In Biology, Controls and Models of Tree Volatile Organic Compound Emissions; Niinemets, Ü., Monson, R.K., Eds.; Tree Physiology; Springer: Dordrecht, The Netherlands, 2013; pp. 451–487. ISBN 978-94-007-6606-8. [Google Scholar]
  97. Müller, J.-F.; Stavrakou, T.; Wallens, S.; De Smedt, I.; Van Roozendael, M.; Potosnak, J.; Rinne, J.; Munger, B.; Goldstein, A.; Guenther, B. Global isoprene emissions estimated using MEGAN, ECMWF analyses and a detailed canopy environment model. Atmos. Chem. Phys. 2008, 8, 1329–1341. [Google Scholar] [CrossRef]
  98. Chi, Y.Q.; Xie, S.D. Spatiotemporal Inventory of Biogenic Volatile Organic Compound Emissions in China Based on Vegetation Volume and Production. Adv. Mater. Res. 2012, 356–360, 2579–2582. [Google Scholar] [CrossRef]
  99. Wang, L.; Lun, X.; Wu, J.; Wang, Q.; Tao, J.; Dou, X.; Zhang, Z. Investigation of biogenic volatile organic compounds emissions in the Qinghai-Tibetan Plateau. Sci. Total Environ. 2023, 902, 165877. [Google Scholar] [CrossRef]
  100. Pierce, T.; Geron, C.; Bender, L.; Dennis, R.; Tonnesen, G.; Guenther, A. Influence of increased isoprene emissions on regional ozone modeling. J. Geophys. Res. Atmos. 1998, 103, 25611–25629. [Google Scholar]
  101. Feldman, M.S.; Howard, T.; McDonald-Buller, E.; Mullins, G.; Allen, D.T.; Hansel, A.; Wisthaler, A. Applications of satellite remote sensing data for estimating biogenic emissions in southeastern Texas. Atmos. Environ. 2010, 44, 917–929. [Google Scholar] [CrossRef]
  102. Zheng, J.; Zheng, Z.; Yu, Y.; Zhong, L. Temporal, spatial characteristics and uncertainty of biogenic VOC emissions in the Pearl River Delta region, China. Atmos. Environ. 2010, 44, 1960–1969. [Google Scholar] [CrossRef]
  103. Sakulyanontvittaya, T.; Duhl, T.; Wiedinmyer, C.; Helmig, D.; Matsunaga, S.; Potosnak, M.; Milford, J.; Guenther, A. Monoterpene and Sesquiterpene Emission Estimates for the United States. Environ. Sci. Technol. 2008, 42, 1623–1629. [Google Scholar] [CrossRef]
  104. Guenther, A.B.; Jiang, X.; Heald, C.L.; Sakulyanontvittaya, T.; Duhl, T.; Emmons, L.K.; Wang, X. The Model of Emissions of Gases and Aerosols from Nature version 2.1 (MEGAN2.1): An extended and updated framework for modeling biogenic emissions. Geosci. Model Dev. 2012, 5, 1471–1492. [Google Scholar] [CrossRef]
  105. Guenther, A.; Jiang, X.; Shah, T.; Huang, L.; Kemball-Cook, S.; Yarwood, G. Model of Emissions of Gases and Aerosol from Nature Version 3 (MEGAN3) for Estimating Biogenic Emissions. In Air Pollution Modeling and its Application XXVI, Proceedings of the ITM 2018, Ottawa, ON, Canada, 14–18 May 2018; Mensink, C., Gong, W., Hakami, A., Eds.; Springer International Publishing: Cham, Switzerland, 2019; pp. 187–192. [Google Scholar]
  106. Sindelarova, K.; Granier, C.; Bouarar, I.; Guenther, A.; Tilmes, S.; Stavrakou, T.; Müller, J.-F.; Kuhn, U.; Stefani, P.; Knorr, W. Global data set of biogenic VOC emissions calculated by the MEGAN model over the last 30 years. Atmos. Chem. Phys. 2014, 14, 9317–9341. [Google Scholar] [CrossRef]
  107. Wang, P.; Schade, G.; Estes, M.; Ying, Q. Improved MEGAN predictions of biogenic isoprene in the contiguous United States. Atmos. Environ. 2017, 148, 337–351. [Google Scholar] [CrossRef]
  108. Jiang, X.; Guenther, A.; Potosnak, M.; Geron, C.; Seco, R.; Karl, T.; Kim, S.; Gu, L.; Pallardy, S. Isoprene emission response to drought and the impact on global atmospheric chemistry. Atmos. Environ. 2018, 183, 69–83. [Google Scholar] [CrossRef] [PubMed]
  109. Bai, J.; Duhl, T. A primary generalized empirical model of BVOC emissions for some typical forests in China. Atmos. Pollut. Res. 2021, 12, 101126. [Google Scholar] [CrossRef]
  110. Benjamin, M.T.; Winer, A.M. Estimating the ozone-forming potential of urban trees and shrubs. Atmos. Environ. 1998, 32, 53–68. [Google Scholar] [CrossRef]
  111. Churkina, G.; Kuik, F.; Bonn, B.; Lauer, A.; Grote, R.; Tomiak, K.; Butler, T.M. Effect of VOC Emissions from Vegetation on Air Quality in Berlin during a Heatwave. Environ. Sci. Technol. 2017, 51, 6120–6130. [Google Scholar] [CrossRef]
  112. Bao, X.; Zhou, W.; Wang, W.; Yao, Y.; Xu, L. Tree species classification improves the estimation of BVOCs from urban greenspace. Sci. Total Environ. 2024, 914, 169762. [Google Scholar] [CrossRef]
  113. Antonelli, M.; Donelli, D.; Barbieri, G.; Valussi, M.; Maggini, V.; Firenzuoli, F. Forest Volatile Organic Compounds and Their Effects on Human Health: A State-of-the-Art Review. Int. J. Environ. Res. Public. Health 2020, 17, 6506. [Google Scholar] [CrossRef]
  114. Meneguzzo, F.; Albanese, L.; Bartolini, G.; Zabini, F. Temporal and Spatial Variability of Volatile Organic Compounds in the Forest Atmosphere. Int. J. Environ. Res. Public. Health 2019, 16, 4915. [Google Scholar] [CrossRef]
  115. Smiatek, G.; Steinbrecher, R. Temporal and spatial variation of forest VOC emissions in Germany in the decade 1994–2003. Atmos. Environ. 2006, 40, 166–177. [Google Scholar] [CrossRef]
  116. Liu, Y.; Schallhart, S.; Tykkä, T.; Räsänen, M.; Merbold, L.; Hellén, H.; Pellikka, P. Biogenic volatile organic compounds in different ecosystems in Southern Kenya. Atmos. Environ. 2021, 246, 118064. [Google Scholar] [CrossRef]
  117. Chang, J.; Ren, Y.; Shi, Y.; Zhu, Y.; Ge, Y.; Hong, S.; Jiao, L.; Lin, F.; Peng, C.; Mochizuki, T.; et al. An inventory of biogenic volatile organic compounds for a subtropical urban-rural complex. Atmos. Environ. 2012, 56, 115–123. [Google Scholar] [CrossRef]
  118. Okumura, M.; Kosugi, Y.; Tani, A. Biogenic volatile organic compound emissions from bamboo species in Japan. J. Agric. Meteorol. 2018, 74, 40–44. [Google Scholar] [CrossRef]
  119. Chen, W.; Wang, L.; Wu, J.; Lun, X.; Wang, X.; Li, X. Emission Pattern of Biogenic Volatile Organic Compounds from Wetland Vegetation. Atmosphere 2024, 15, 651. [Google Scholar] [CrossRef]
  120. Mannisto, E.; Ylanne, H.; Losoi, M.; Keinanen, M.; Yli-Pirila, P.; Korrensalo, A.; Back, J.; Hellen, H.; Virtanen, A.; Tuittila, E.-S. Emissions of biogenic volatile organic compounds from adjacent boreal fen and bog as impacted by vegetation composition. Sci. Total Environ. 2023, 858, 159809. [Google Scholar] [CrossRef]
  121. Gomez, L.G.; Loubet, B.; Lafouge, F.; Ciuraru, R.; Buysse, P.; Durand, B.; Gueudet, J.-C.; Fanucci, O.; Fortineau, A.; Zurfluh, O.; et al. Comparative study of biogenic volatile organic compounds fluxes by wheat, maize and rapeseed with dynamic chambers over a short period in northern France. Atmos. Environ. 2019, 214, 116855. [Google Scholar] [CrossRef]
  122. Mozaffar, A.; Schoon, N.; Digrado, A.; Bachy, A.; Delaplace, P.; du Jardin, P.; Fauconnier, M.-L.; Aubinet, M.; Heinesch, B.; Amelynck, C. Methanol emissions from maize: Ontogenetic dependence to varying light conditions and guttation as an additional factor constraining the flux. Atmos. Environ. 2017, 152, 405–417. [Google Scholar] [CrossRef]
  123. Bäck, J.; Aaltonen, H.; Hellén, H.; Kajos, M.K.; Patokoski, J.; Taipale, R.; Pumpanen, J.; Heinonsalo, J. Variable emissions of microbial volatile organic compounds (MVOCs) from root-associated fungi isolated from Scots pine. Atmos. Environ. 2010, 44, 3651–3659. [Google Scholar] [CrossRef]
  124. Bourtsoukidis, E.; Behrendt, T.; Yañez-Serrano, A.M.; Hellén, H.; Diamantopoulos, E.; Catão, E.; Ashworth, K.; Pozzer, A.; Quesada, C.A.; Martins, D.L.; et al. Strong sesquiterpene emissions from Amazonian soils. Nat. Commun. 2018, 9, 2226. [Google Scholar] [CrossRef]
  125. Mäki, M. Volatile Organic Compound Fluxes from Northern Forest Soils. Ph.D. Thesis, Institute for Atmospheric and Earth System Research, Forest Sciences Faculty of Agriculture and Forestry, University of Helsinki, Helsinki, Finland, 2019. [Google Scholar]
  126. Peñuelas, J.; Asensio, D.; Tholl, D.; Wenke, K.; Rosenkranz, M.; Piechulla, B.; Schnitzler, J.P. Biogenic volatile emissions from the soil. Plant Cell Environ. 2014, 37, 1866–1891. [Google Scholar] [CrossRef] [PubMed]
  127. Wester-Larsen, L.; Kramshøj, M.; Albers, C.N.; Rinnan, R. Biogenic Volatile Organic Compounds in Arctic Soil: A Field Study of Concentrations and Variability with Vegetation Cover. J. Geophys. Res. Biogeosci. 2020, 125, e2019JG005551. [Google Scholar] [CrossRef]
  128. Sindelarova, K.; Markova, J.; Simpson, D.; Huszar, P.; Karlicky, J.; Darras, S.; Granier, C. High-resolution biogenic global emission inventory for the time period 2000–2019 for air quality modelling. Earth Syst. Sci. Data 2022, 14, 251–270. [Google Scholar] [CrossRef]
  129. Yáñez-Serrano, A.M.; Bourtsoukidis, E.; Alves, E.G.; Bauwens, M.; Stavrakou, T.; Llusià, J.; Filella, I.; Guenther, A.; Williams, J.; Artaxo, P.; et al. Amazonian biogenic volatile organic compounds under global change. Global Change Biol. 2020, 26, 4722–4751. [Google Scholar] [CrossRef] [PubMed]
  130. Yee, L.D.; Isaacman-VanWertz, G.; Wernis, R.A.; Meng, M.; Rivera, V.; Kreisberg, N.M.; Hering, S.V.; Bering, M.S.; Glasius, M.; Upshur, M.A.; et al. Observations of sesquiterpenes and their oxidation products in central Amazonia during the wet and dry seasons. Atmos. Chem. Phys. 2018, 18, 10433–10457. [Google Scholar] [CrossRef]
  131. Wang, X.; Zhang, Y.; Tan, Y.; Tan, Y.; Bai, J.; Gu, D.; Ma, Z.; Du, J.; Han, Z. Effects of light on the emissions of biogenic isoprene and monoterpenes: A review. Atmos. Pollut. Res. 2022, 13, 101397. [Google Scholar] [CrossRef]
  132. Liu, L.; Seyler, B.C.; Liu, H.; Zhou, L.; Chen, D.; Liu, S.; Yan, C.; Yang, F.; Song, D.; Tan, Q.; et al. Biogenic volatile organic compound emission patterns and secondary pollutant formation potentials of dominant greening trees in Chengdu, southwest China. J. Environ. Sci. 2022, 114, 179–193. [Google Scholar] [CrossRef]
  133. Cheng, X.; Li, H.; Zhang, Y.; Li, Y.; Zhang, W.; Wang, X.; Bi, F.; Zhang, H.; Gao, J.; Chai, F.; et al. Atmospheric isoprene and monoterpenes in a typical urban area of Beijing: Pollution characterization, chemical reactivity and source identification. J. Environ. Sci. 2018, 71, 150–167. [Google Scholar] [CrossRef]
  134. Liu, J.; Chu, B.; Chen, T.; Liu, C.; Wang, L.; Bao, X.; He, H. Secondary Organic Aerosol Formation from Ambient Air at an Urban Site in Beijing: Effects of OH Exposure and Precursor Concentrations. Environ. Sci. Technol. 2018, 52, 6834–6841. [Google Scholar] [CrossRef]
  135. Ndah, F.; Valolahti, H.; Schollert, M.; Michelsen, A.; Rinnan, R.; Kivimäenpää, M. Influence of increased nutrient availability on biogenic volatile organic compound (BVOC) emissions and leaf anatomy of subarctic dwarf shrubs under climate warming and increased cloudiness. Ann. Bot. 2022, 129, 443–455. [Google Scholar] [CrossRef]
  136. Lugo, P.L.; Straccia, V.G.; Rivela, C.B.; Patroescu-Klotz, I.; Illmann, N.; Teruel, M.A.; Wiesen, P.; Blanco, M.B. Diurnal photodegradation of fluorinated diketones (FDKs) by OH radicals using different atmospheric simulation chambers: Role of keto-enol tautomerization on reactivity. Chemosphere 2022, 286, 131562. [Google Scholar] [CrossRef] [PubMed]
  137. Frazier, G.; McGlynn, D.F.; Barry, L.E.; Lerdau, M.; Pusede, S.E.; Isaacman-VanWertz, G. Composition, concentration, and oxidant reactivity of sesquiterpenes in the southeastern U.S. Environ. Sci. Atmos. 2022, 2, 1208–1220. [Google Scholar] [CrossRef]
  138. McGlynn, D.F.; Barry, L.E.R.; Lerdau, M.T.; Pusede, S.E.; Isaacman-VanWertz, G. Measurement report: Variability in the composition of biogenic volatile organic compounds in a Southeastern US forest and their role in atmospheric reactivity. Atmos. Chem. Phys. 2021, 21, 15755–15770. [Google Scholar] [CrossRef]
  139. Atkinson, R.; Arey, J. Gas-phase tropospheric chemistry of biogenic volatile organic compounds: A review. Atmos. Environ. 2003, 37, 197–219. [Google Scholar] [CrossRef]
  140. Bouvier-Brown, N.C.; Goldstein, A.H.; Gilman, J.B.; Kuster, W.C.; de Gouw, J.A. In-situ ambient quantification of monoterpenes, sesquiterpenes, and related oxygenated compounds during BEARPEX 2007: Implications for gas- and particle-phase chemistry. Atmos. Chem. Phys. 2009, 9, 5505–5518. [Google Scholar] [CrossRef]
  141. Calvert, J.; Mellouki, A.; Orlando, J.; Pilling, M.; Wallington, T. Mechanisms of Atmospheric Oxidation of the Oxygenates; Oxford University Press: Oxford, UK, 2011. [Google Scholar]
  142. Isidorov, V.A.; Masłowiecka, J. Chemical Composition of Volatile and Extractive Organic Compounds in the Inflorescence Litter of Five Species of Woody Plants. Plants 2024, 13, 1829. [Google Scholar] [CrossRef]
  143. Sharkey, T.D.; Wiberley, A.E.; Donohue, A.R. Isoprene Emission from Plants: Why and How. Ann. Bot. 2008, 101, 5–18. [Google Scholar] [CrossRef]
  144. Ma, S.X.; Rindelaub, J.D.; McAvey, K.M.; Gagare, P.D.; Nault, B.A.; Ramachandran, P.V.; Shepson, P.B. α-Pinene nitrates: Synthesis, yields and atmospheric chemistry. Atmos. Chem. Phys. 2011, 11, 6337–6347. [Google Scholar] [CrossRef]
  145. Dash, M.R.; Rajakumar, B. Experimental and theoretical rate coefficients for the gas phase reaction of β-Pinene with OH radical. Atmos. Environ. 2013, 79, 161–171. [Google Scholar] [CrossRef]
  146. Chen, J.; Griffin, R.J. Modeling secondary organic aerosol formation from oxidation of α-pinene, β-pinene, and d-limonene. Atmos. Environ. 2005, 39, 7731–7744. [Google Scholar] [CrossRef]
  147. Chen, X.; Hopke, P.K. A chamber study of secondary organic aerosol formation by linalool ozonolysis. Atmos. Environ. 2009, 43, 3935–3940. [Google Scholar] [CrossRef]
  148. Arneth, A.; Schurgers, G.; Lathiere, J.; Duhl, T.; Beerling, D.J.; Hewitt, C.N.; Martin, M.; Guenther, A. Global terrestrial isoprene emission models: Sensitivity to variability in climate and vegetation. Atmos. Chem. Phys. 2011, 11, 8037–8052. [Google Scholar] [CrossRef]
  149. Cao, Y.; Yue, X.; Liao, H.; Yang, Y.; Zhu, J.; Chen, L.; Tian, C.; Lei, Y.; Zhou, H.; Ma, Y. Ensemble projection of global isoprene emissions by the end of 21st century using CMIP6 models. Atmos. Environ. 2021, 267, 118766. [Google Scholar] [CrossRef]
  150. Weng, H.; Lin, J.; Martin, R.; Millet, D.B.; Jaeglé, L.; Ridley, D.; Keller, C.; Li, C.; Du, M.; Meng, J. Global high-resolution emissions of soil NOx, sea salt aerosols, and biogenic volatile organic compounds. Sci. Data 2020, 7, 148. [Google Scholar] [CrossRef]
  151. Rabin, S.S.; Alexander, P.; Henry, R.; Anthoni, P.; Pugh, T.A.M.; Rounsevell, M.; Arneth, A. Impacts of future agricultural change on ecosystem service indicators. Earth Syst. Dyn. 2020, 11, 357–376. [Google Scholar] [CrossRef]
  152. Sancho-Knapik, D.; Gil-Pelegrin, E.; Pedro Ferrio, J.; Alonso-Forn, D.; Martin-Sanchez, R.; dos Santos Silva, J.V.; Imanishi, J.; Javier Peguero-Pina, J.; Angeles Sanz, M. Changes in the Abundance of Monoterpenes from Breathable Air of a Mediterranean Conifer Forest: When Is the Best Time for a Human Healthy Leisure Activity? Forests 2022, 13, 965. [Google Scholar] [CrossRef]
  153. Acosta Navarro, J.C.; Smolander, S.; Struthers, H.; Zorita, E.; Ekman, A.M.L.; Kaplan, J.O.; Guenther, A.; Arneth, A.; Riipinen, I. Global emissions of terpenoid VOCs from terrestrial vegetation in the last millennium. J. Geophys. Res. Atmos. 2014, 119, 6867–6885. [Google Scholar] [CrossRef]
  154. Llusia, J.; Penuelas, J.; Guenther, A.; Rapparini, F. Seasonal variations in terpene emission factors of dominant species in four ecosystems in NE Spain. Atmos. Environ. 2013, 70, 149–158. [Google Scholar] [CrossRef]
  155. Wang, H.; Liu, X.; Wu, C.; Lin, G. Regional to global distributions, trends, and drivers of biogenic volatile organic compound emission from 2001 to 2020. Atmos. Chem. Phys. 2024, 24, 3309–3328. [Google Scholar] [CrossRef]
  156. Dudareva, N.; Negre, F.; Nagegowda, D.A.; Orlova, I. Plant Volatiles: Recent Advances and Future Perspectives. Crit. Rev. Plant Sci. 2006, 25, 417–440. [Google Scholar] [CrossRef]
  157. Lathière, J.; Hauglustaine, D.A.; Friend, A.D.; De Noblet-Ducoudré, N.; Viovy, N.; Folberth, G.A. Impact of climate variability and land use changes on global biogenic volatile organic compound emissions. Atmos. Chem. Phys. 2006, 6, 2129–2146. [Google Scholar] [CrossRef]
  158. Xin-min, Z.; Fa-he, C.; Ting-ting, Y.; Kai, Z.; Yi-zhen, C.; Jing-chun, D.; Zhi-gang, X. Photochemical Characteristics and Sources of Volatile Organic Compounds in Wuqing, Tianjin. Res. Environ. Sci. 2012, 25, 1085–1091. [Google Scholar]
  159. Zhu, S.; Huang, X.; He, L.; Lu, S.; Feng, N. Variation characteristics and chemical reactivity of ambient VOCs in Shenzhen. China Environ. Sci. 2012, 32, 2140–2148. [Google Scholar]
  160. Zou, Y.; Deng, X.; Li, F.; Wang, B.; Tan, H.; Deng, T.; Mai, B.; Liu, X. Variation characteristics, chemical reactivity and sources of isoprene in the atmosphere of Guangzhou. Huanjing Kexue Xuebao 2015, 35, 647–655. [Google Scholar]
  161. Ran, L.; Zhao, C.; Geng, F.; Tie, X.; Tang, X.; Peng, L.; Zhou, G.; Yu, Q.; Xu, J.; Guenther, A. Ozone photochemical production in urban Shanghai, China: Analysis based on ground level observations. J. Geophys. Res. 2009, 114. [Google Scholar] [CrossRef]
  162. Park, C.; Schade, G.W.; Boedeker, I. Characteristics of the flux of isoprene and its oxidation products in an urban area. J. Geophys. Res. Atmos. 2011, 116. [Google Scholar] [CrossRef]
  163. Zhang, J.-K.; Wang, Y.-S.; Wu, F.-K.; Sun, J. Study on atmospheric VOCs in Gongga Mountain base station. Huanjing Kexue 2012, 33, 4159–4166. [Google Scholar]
  164. Ying-xin, Y.U.; Sheng, W.E.N.; Hui-xiong, L.V.; Xin-ming, W.; Guo-ying, S.; Jia-mo, F.U. Diurnal Variations of VOCs and Relative Contribution to Ozone in Forest of Guangzhou. Huanjing Kexue Yu Jishu 2009, 32, 94–98. [Google Scholar]
  165. Dudareva, N.; Klempien, A.; Muhlemann, J.K.; Kaplan, I. Biosynthesis, function and metabolic engineering of plant volatile organic compounds. New Phytol. 2013, 198, 16–32. [Google Scholar] [CrossRef]
  166. Li, Z.; Sharkey, T.D. Molecular and Pathway Controls on Biogenic Volatile Organic Compound Emissions. In Biology, Controls and Models of Tree Volatile Organic Compound Emissions; Niinemets, Ü., Monson, R.K., Eds.; Springer: Dordrecht, The Netherlands, 2013; pp. 119–151. ISBN 978-94-007-6606-8. [Google Scholar]
  167. Ma, M.; Li, M.; Wu, Z.; Liang, X.; Zheng, Q.; Li, D.; Wang, G.; An, T. The microbial biosynthesis of noncanonical terpenoids. Appl. Microbiol. Biotechnol. 2024, 108, 226. [Google Scholar] [CrossRef]
  168. Tholl, D. Biosynthesis and Biological Functions of Terpenoids in Plants. In Biotechnology of Isoprenoids; Schrader, J., Bohlmann, J., Eds.; Springer International Publishing: Cham, Switzerland, 2015; pp. 63–106. ISBN 978-3-319-20107-8. [Google Scholar]
  169. Arigoni, D.; Sagner, S.; Latzel, C.; Eisenreich, W.; Bacher, A.; Zenk, M.H. Terpenoid biosynthesis from 1-deoxy-D-xylulose in higher plants by intramolecular skeletal rearrangement. Proc. Natl. Acad. Sci. USA 1997, 94, 10600–10605. [Google Scholar] [CrossRef] [PubMed]
  170. Bongers, M.; Perez-Gil, J.; Hodson, M.P.; Schrübbers, L.; Wulff, T.; Sommer, M.O.; Nielsen, L.K.; Vickers, C.E. Adaptation of hydroxymethylbutenyl diphosphate reductase enables volatile isoprenoid production. eLife 2020, 9, e48685. [Google Scholar] [CrossRef] [PubMed]
  171. Niinemets, Ü.; Sun, Z.; Talts, E. Controls of the quantum yield and saturation light of isoprene emission in different-aged aspen leaves. Plant Cell Environ. 2015, 38, 2707–2720. [Google Scholar] [CrossRef]
  172. Sun, Z.; Shen, Y.; Niinemets, Ü. Responses of isoprene emission and photochemical efficiency to severe drought combined with prolonged hot weather in hybrid Populus. J. Exp. Bot. 2020, 71, 7364–7381. [Google Scholar] [CrossRef] [PubMed]
  173. Tritsch, D.; Hemmerlin, A.; Bach, T.J.; Rohmer, M. Plant isoprenoid biosynthesis via the MEP pathway: In vivo IPP/DMAPP ratio produced by (E)-4-hydroxy-3-methylbut-2-enyl diphosphate reductase in tobacco BY-2 cell cultures. FEBS Lett. 2010, 584, 129–134. [Google Scholar] [CrossRef]
  174. Rodríguez-Concepción, M.; Boronat, A. Breaking new ground in the regulation of the early steps of plant isoprenoid biosynthesis. Curr. Opin. Plant Biol. 2015, 25, 17–22. [Google Scholar] [CrossRef]
  175. Liao, P.; Hemmerlin, A.; Bach, T.J.; Chye, M.-L. The potential of the mevalonate pathway for enhanced isoprenoid production. Biotechnol. Adv. 2016, 34, 697–713. [Google Scholar] [CrossRef]
  176. Bergman, M.E.; Kortbeek, R.W.J.; Gutensohn, M.; Dudareva, N. Plant terpenoid biosynthetic network and its multiple layers of regulation. Prog. Lipid Res. 2024, 95, 101287. [Google Scholar] [CrossRef]
  177. Lange, B.M.; Ahkami, A. Metabolic engineering of plant monoterpenes, sesquiterpenes and diterpenes—Current status and future opportunities. Plant Biotechnol. J. 2013, 11, 169–196. [Google Scholar] [CrossRef]
  178. Liu, B.; Zhang, L.; Rusalepp, L.; Kaurilind, E.; Sulaiman, H.Y.; Püssa, T.; Niinemets, Ü. Heat priming improved heat tolerance of photosynthesis, enhanced terpenoid and benzenoid emission and phenolics accumulation in. Plant Cell Environ. 2021, 44, 2365–2385. [Google Scholar] [CrossRef]
  179. Vellosillo, T.; Martínez, M.; López, M.A.; Vicente, J.; Cascón, T.; Dolan, L.; Hamberg, M.; Castresana, C. Oxylipins Produced by the 9-Lipoxygenase Pathway in Arabidopsis Regulate Lateral Root Development and Defense Responses through a Specific Signaling Cascade. The Plant Cell 2007, 19, 831–846. [Google Scholar] [PubMed]
  180. Wan, S.; Xin, X.-F. Regulation and integration of plant jasmonate signaling: A comparative view of monocot and dicot. J. Genet. Genom. 2022, 49, 704–714. [Google Scholar] [CrossRef]
  181. Wang, M.; Fan, X.; Ding, F. Jasmonate: A Hormone of Primary Importance for Temperature Stress Response in Plants. Plants 2023, 12, 4080. [Google Scholar] [CrossRef] [PubMed]
  182. Guenther, A.; Geron, C.; Pierce, T.; Lamb, B.; Harley, P.; Fall, R. Natural emissions of non-methane volatile organic compounds, carbon monoxide, and oxides of nitrogen from North America. Atmos. Environ. 2000, 34, 2205–2230. [Google Scholar] [CrossRef]
  183. Potosnak, M.J.; LeStourgeon, L.; Pallardy, S.G.; Hosman, K.P.; Gu, L.; Karl, T.; Geron, C.; Guenther, A.B. Observed and modeled ecosystem isoprene fluxes from an oak-dominated temperate forest and the influence of drought stress. Atmos. Environ. 2014, 84, 314–322. [Google Scholar] [CrossRef]
  184. Yang, W.; Zhang, B.; Wu, Y.; Liu, S.; Kong, F.; Li, L. Effects of soil drought and nitrogen deposition on BVOC emissions and their O3 and SOA formation for Pinus thunbergii. Environ. Pollut. 2023, 316, 120693. [Google Scholar] [CrossRef] [PubMed]
  185. Peng, Y.; Mouat, A.P.; Hu, Y.; Li, M.; McDonald, B.C.; Kaiser, J. Source appointment of volatile organic compounds and evaluation of anthropogenic monoterpene emission estimates in Atlanta, Georgia. Atmos. Environ. 2022, 288, 119324. [Google Scholar]
  186. Xu, L.; Pye, H.O.T.; He, J.; Chen, Y.; Murphy, B.N.; Ng, N.L. Experimental and model estimates of the contributions from biogenic monoterpenes and sesquiterpenes to secondary organic aerosol in the southeastern United States. Atmos. Chem. Phys. 2018, 18, 12613–12637. [Google Scholar] [CrossRef]
  187. Loreto, F.; Velikova, V. Isoprene Produced by Leaves Protects the Photosynthetic Apparatus against Ozone Damage, Quenches Ozone Products, and Reduces Lipid Peroxidation of Cellular Membranes. Plant Physiol. 2001, 127, 1781–1787. [Google Scholar] [CrossRef]
  188. Zenone, T.; Hendriks, C.; Brilli, F.; Fransen, E.; Gioli, B.; Portillo-Estrada, M.; Schaap, M.; Ceulemans, R. Interaction between isoprene and ozone fluxes in a poplar plantation and its impact on air quality at the European level. Sci. Rep. 2016, 6, 32676. [Google Scholar] [CrossRef]
  189. Zou, Y.; Deng, X.J.; Deng, T.; Yin, C.Q.; Li, F. One-Year Characterization and Reactivity of Isoprene and Its Impact on Surface Ozone Formation at A Suburban Site in Guangzhou, China. Atmosphere 2019, 10, 201. [Google Scholar] [CrossRef]
  190. Schade, G.W.; Goldstein, A.H.; Lamanna, M.S. Are monoterpene emissions influenced by humidity? Geophys. Res. Lett. 1999, 26, 2187–2190. [Google Scholar] [CrossRef]
  191. Street, R.A.; Owen, S.; Duckham, S.C.; Boissard, C.; Hewitt, C.N. Effect of habitat and age on variations in volatile organic compound (VOC) emissions from Quercus ilex and Pinus pinea. Atmos. Environ. 1997, 31, 89–100. [Google Scholar] [CrossRef]
  192. Kuzma, J.; Fall, R. Leaf Isoprene Emission Rate Is Dependent on Leaf Development and the Level of Isoprene Synthase. Plant Physiol. 1993, 101, 435–440. [Google Scholar] [CrossRef] [PubMed]
  193. Rasulov, B.; Bichele, I.; Laisk, A.; Niinemets, Ü. Competition between isoprene emission and pigment synthesis during leaf development in aspen. Plant Cell Environ. 2014, 37, 724–741. [Google Scholar] [CrossRef]
  194. Lin, P.-A.; Chen, Y.; Ponce, G.; Acevedo, F.E.; Lynch, J.P.; Anderson, C.T.; Ali, J.G.; Felton, G.W. Stomata-mediated interactions between plants, herbivores, and the environment. Trends Plant Sci. 2022, 27, 287–300. [Google Scholar] [CrossRef]
  195. Loreto, F.; Schnitzler, J.-P. Abiotic stresses and induced BVOCs. Trends Plant Sci. 2010, 15, 154–166. [Google Scholar] [CrossRef] [PubMed]
  196. Niinemets, Ü.; Reichstein, M. Controls on the emission of plant volatiles through stomata: A sensitivity analysis. J. Geophys. Res. Atmos. 2003, 108, 4211. [Google Scholar] [CrossRef]
  197. Yin, Q.; Tian, T.; Kou, M.; Liu, P.; Wang, L.; Hao, Z.; Yue, M. The relationships between photosynthesis and stomatal traits on the Loess Plateau. Glob. Ecol. Conserv. 2020, 23, e01146. [Google Scholar] [CrossRef]
  198. Chuong, B.; Zhang, J.; Donahue, N.M. Cycloalkene Ozonolysis: Collisionally Mediated Mechanistic Branching. J. Am. Chem. Soc. 2004, 126, 12363–12373. [Google Scholar] [CrossRef]
  199. Nemecek-Marshall, M.; MacDonald, R.C.; Franzen, J.J.; Wojciechowski, C.L.; Fall, R. Methanol Emission from Leaves (Enzymatic Detection of Gas-Phase Methanol and Relation of Methanol Fluxes to Stomatal Conductance and Leaf Development). Plant Physiol. 1995, 108, 1359–1368. [Google Scholar] [CrossRef] [PubMed]
  200. Niinemets, Ü. Mild versus severe stress and BVOCs: Thresholds, priming and consequences. Trends Plant Sci. 2010, 15, 145–153. [Google Scholar] [CrossRef]
  201. Riedlmeier, M.; Ghirardo, A.; Wenig, M.; Knappe, C.; Koch, K.; Georgii, E.; Dey, S.; Parker, J.E.; Schnitzler, J.-P.; Vlot, A.C. Monoterpenes Support Systemic Acquired Resistance within and between Plants. Plant Cell 2017, 29, 1440–1459. [Google Scholar] [CrossRef] [PubMed]
  202. Imbiscuso, G.; Trotta, A.; Maffei, M.; Bossi, S. Herbivory induces a ROS burst and the release of volatile organic compounds in the fern Pteris vittata L. J. Plant Interact. 2009, 4, 15–22. [Google Scholar] [CrossRef]
  203. Joutsensaari, J.; Yli-Pirilä, P.; Korhonen, H.; Arola, A.; Blande, J.D.; Heijari, J.; Kivimäenpää, M.; Mikkonen, S.; Hao, L.; Miettinen, P.; et al. Biotic stress accelerates formation of climate-relevant aerosols in boreal forests. Atmos. Chem. Phys. 2015, 15, 12139–12157. [Google Scholar] [CrossRef]
  204. Penuelas, J.; Llusia, J. The Complexity of Factors Driving Volatile Organic Compound Emissions by Plants. Biol. Plant. 2001, 44, 481–487. [Google Scholar] [CrossRef]
  205. Litvak, M.E.; Monson, R.K. Patterns of induced and constitutive monoterpene production in conifer needles in relation to insect herbivory. Oecologia 1998, 114, 531–540. [Google Scholar] [CrossRef]
  206. Abis, L.; Kalalian, C.; Lunardelli, B.; Wang, T.; Zhang, L.; Chen, J.; Perrier, S.; Loubet, B.; Ciuraru, R.; George, C. Measurement report: Biogenic volatile organic compound emission profiles of rapeseed leaf litter and its secondary organic aerosol formation potential. Atmos. Chem. Phys. 2021, 21, 12613–12629. [Google Scholar] [CrossRef]
  207. Monson, R.K.; Jaeger, C.H.; Adams, W.W.; Driggers, E.M.; Silver, G.M.; Fall, R. Relationships among Isoprene Emission Rate, Photosynthesis, and Isoprene Synthase Activity as Influenced by Temperature. Plant Physiol. 1992, 98, 1175–1180. [Google Scholar] [CrossRef]
  208. Li, T.; Holst, T.; Michelsen, A.; Rinnan, R. Amplification of plant volatile defence against insect herbivory in a warming Arctic tundra. Nat. Plants 2019, 5, 568–574. [Google Scholar] [CrossRef]
  209. Loreto, F.; Barta, C.; Brilli, F.; Nogues, I. On the induction of volatile organic compound emissions by plants as consequence of wounding or fluctuations of light and temperature. Plant Cell Environ. 2006, 29, 1820–1828. [Google Scholar] [CrossRef] [PubMed]
  210. Lerdau, M.; Gray, D. Ecology and evolution of light-dependent and light-independent phytogenic volatile organic carbon. New Phytol. 2003, 157, 199–211. [Google Scholar] [CrossRef]
  211. Sharkey, T.D.; Monson, R.K. Isoprene research—60 years later, the biology is still enigmatic. Plant Cell Environ. 2017, 40, 1671–1678. [Google Scholar] [CrossRef]
  212. Yu, S.-J.; Seo, Y.-S.; Son, Y.-S.; Kim, J.-C. Characteristics of BVOCs emitted from representative tree species in South Korea according to environmental influences. Atmos. Environ. 2024, 318, 120238. [Google Scholar] [CrossRef]
  213. Aliniaeifard, S.; van Meeteren, U. Can prolonged exposure to low VPD disturb the ABA signalling in stomatal guard cells? J. Exp. Bot. 2013, 64, 3551–3566. [Google Scholar] [CrossRef]
  214. Arve, L.E.; Terfa, M.T.; Gislerød, H.R.; Olsen, J.E.; Torre, S. High relative air humidity and continuous light reduce stomata functionality by affecting the ABA regulation in rose leaves. Plant Cell Environ. 2013, 36, 382–392. [Google Scholar] [CrossRef] [PubMed]
  215. Mu, Z.; Asensio, D.; Llusià, J.; Filella, I.; Ogaya, R.; Yi, Z.; Peñuelas, J. Annual and seasonal variations in soil volatile organic compound concentrations in a Mediterranean shrubland and holm oak forest. Geoderma 2022, 405, 115401. [Google Scholar] [CrossRef]
  216. Yáñez-Serrano, A.M.; Nölscher, A.C.; Williams, J.; Wolff, S.; Alves, E.; Martins, G.A.; Bourtsoukidis, E.; Brito, J.; Jardine, K.; Artaxo, P.; et al. Diel and seasonal changes of biogenic volatile organic compounds within and above an Amazonian rainforest. Atmos. Chem. Phys. 2015, 15, 3359–3378. [Google Scholar] [CrossRef]
  217. Fanourakis, D.; Giday, H.; Hyldgaard, B.; Bouranis, D.; Körner, O.; Ottosen, C.-O. Low air humidity during cultivation promotes stomatal closure ability in rose. Eur. J. Hortic. Sci. 2019, 84, 245–252. [Google Scholar] [CrossRef]
  218. Ning, P.; Guo, X.; Tian, S.; Shi, J.; Sun, C. Emission of main BVOCS for typical landscape trees in Kunming. Zhongnan Daxue Xuebao (Ziran Kexue Ban)/J. Cent. South Univ. (Sci. Technol.) 2013, 44, 1290–1296. [Google Scholar]
  219. Constable, J.V.; Litvak, M.E.; Greenberg, J.P.; Monson, R.K. Monoterpene emission from coniferous trees in response to elevated CO2 concentration and climate warming. Glob. Change Biol. 1999, 5, 252–267. [Google Scholar] [CrossRef]
  220. Jasoni, R.; Kane, C.; Green, C.; Peffley, E.; Tissue, D.; Thompson, L.; Payton, P.; Paré, P.W. Altered leaf and root emissions from onion (Allium cepa L.) grown under elevated CO2 conditions. Environ. Exp. Bot. 2004, 51, 273–280. [Google Scholar] [CrossRef]
  221. Rapparini, F.; Baraldi, R.; Miglietta, F.; Loreto, F. Isoprenoid emission in trees of Quercus pubescens and Quercus ilex with lifetime exposure to naturally high CO2 environment. Plant Cell Environ. 2004, 27, 381–391. [Google Scholar] [CrossRef]
  222. Heiden, A.C.; Kobel, K.; Langebartels, C.; Schuh-Thomas, G.; Wildt, J. Emissions of Oxygenated Volatile Organic Compounds from Plants Part I: Emissions from Lipoxygenase Activity. J. Atmos. Chem. 2003, 45, 143–172. [Google Scholar] [CrossRef]
  223. Kim, S.-Y.; Jiang, X.; Lee, M.; Turnipseed, A.; Guenther, A.; Kim, J.-C.; Lee, S.-J.; Kim, S. Impact of biogenic volatile organic compounds on ozone production at the Taehwa Research Forest near Seoul, South Korea. Atmos. Environ. 2013, 70, 447–453. [Google Scholar] [CrossRef]
  224. Sha, T.; Ma, X.; Zhang, H.; Janechek, N.; Wang, Y.; Wang, Y.; Castro García, L.; Jenerette, G.D.; Wang, J. Impacts of Soil NOx Emission on O3 Air Quality in Rural California. Environ. Sci. Technol. 2021, 55, 7113–7122. [Google Scholar] [CrossRef]
  225. Tiwari, A.; Kumar, P.; Baldauf, R.; Zhang, K.M.; Pilla, F.; Di Sabatino, S.; Brattich, E.; Pulvirenti, B. Considerations for evaluating green infrastructure impacts in microscale and macroscale air pollution dispersion models. Sci. Total Environ. 2019, 672, 410–426. [Google Scholar] [CrossRef]
  226. Llusià, J.; Peñuelas, J.; Gimeno, B.S. Seasonal and species-specific response of VOC emissions by Mediterranean woody plant to elevated ozone concentrations. Atmos. Environ. 2002, 36, 3931–3938. [Google Scholar] [CrossRef]
  227. Li, L.; Cao, J.; Hao, Y. Spatial and species-specific responses of biogenic volatile organic compound (BVOC) emissions to elevated ozone from 2014–2020 in China. Sci. Total Environ. 2023, 868, 161636. [Google Scholar] [CrossRef]
  228. Vo, T.; Faiola, C.L. Acute ozone exposure decreases terpene emissions from Canary Island pines. Agric. For. Meteorol. 2023, 333, 109416. [Google Scholar] [CrossRef]
  229. Barreira, L.M.F.; Ylisirniö, A.; Pullinen, I.; Buchholz, A.; Li, Z.; Lipp, H.; Junninen, H.; Hõrrak, U.; Noe, S.M.; Krasnova, A.; et al. The importance of sesquiterpene oxidation products for secondary organic aerosol formation in a springtime hemiboreal forest. Atmos. Chem. Phys. 2021, 21, 11781–11800. [Google Scholar] [CrossRef]
  230. Salvador, C.M.; Chou, C.C.-K.; Ho, T.-T.; Tsai, C.-Y.; Tsao, T.-M.; Tsai, M.-J.; Su, T.-C. Contribution of Terpenes to Ozone Formation and Secondary Organic Aerosols in a Subtropical Forest Impacted by Urban Pollution. Atmosphere 2020, 11, 1232. [Google Scholar] [CrossRef]
  231. Charan, S.M.; Huang, Y.; Seinfeld, J.H. Computational Simulation of Secondary Organic Aerosol Formation in Laboratory Chambers. Chem. Rev. 2019, 119, 11912–11944. [Google Scholar] [CrossRef]
  232. Hellén, H.; Praplan, A.P.; Tykkä, T.; Ylivinkka, I.; Vakkari, V.; Bäck, J.; Petäjä, T.; Kulmala, M.; Hakola, H. Long-term measurements of volatile organic compounds highlight the importance of sesquiterpenes for the atmospheric chemistry of a boreal forest. Atmos. Chem. Phys. 2018, 18, 13839–13863. [Google Scholar] [CrossRef]
  233. Kawamura, K.; Bikkina, S. A review of dicarboxylic acids and related compounds in atmospheric aerosols: Molecular distributions, sources and transformation. Atmos. Res. 2016, 170, 140–160. [Google Scholar] [CrossRef]
  234. Yue, H.; Zhang, C.; Lin, X.; Wen, Z.; Zhang, W.; Mostafa, S.; Luo, P.-L.; Zhang, Z.; Hemberger, P.; Fittschen, C.; et al. Dimeric Product of Peroxy Radical Self-Reaction Probed with VUV Photoionization Mass Spectrometry and Theoretical Calculations: The Case of C2H5OOC2H5. Int. J. Mol. Sci. 2023, 24, 3731. [Google Scholar] [CrossRef] [PubMed]
  235. Dunlea, E.J.; Ravishankara, A.R. Measurement of the rate coefficient for the reaction of O(1D) with H2O and re-evaluation of the atmospheric OH production rate. Phys. Chem. Chem. Phys. 2004, 6, 3333–3340. [Google Scholar] [CrossRef]
  236. Monks, P.S. Gas-phase radical chemistry in the troposphere. Chem. Soc. Rev. 2005, 34, 376–395. [Google Scholar] [CrossRef]
  237. Wang, X.; Liu, L.; Fang, W.; Chen, X. Theoretical insight towards the photo-dissociation dynamics of O3–H2O complex: Deep understanding the source of atmospheric hydroxyl radical. Chem. Phys. Lett. 2014, 608, 95–101. [Google Scholar] [CrossRef]
  238. Atkinson, R. Kinetics and mechanisms of the gas-phase reactions of the hydroxyl radical with organic compounds under atmospheric conditions. Chem. Rev. 1986, 86, 69–201. [Google Scholar] [CrossRef]
  239. Zhao, Y.; Wingen, L.M.; Perraud, V.; Greaves, J.; Finlayson-Pitts, B.J. Role of the reaction of stabilized Criegee intermediates with peroxy radicals in particle formation and growth in air. Phys. Chem. Chem. Phys. 2015, 17, 12500–12514. [Google Scholar] [CrossRef] [PubMed]
  240. Brown, S.S.; Stutz, J. Nighttime radical observations and chemistry. Chem. Soc. Rev. 2012, 41, 6405–6447. [Google Scholar] [CrossRef] [PubMed]
  241. Geyer, A.; Alicke, B.; Konrad, S.; Schmitz, T.; Stutz, J.; Platt, U. Chemistry and oxidation capacity of the nitrate radical in the continental boundary layer near Berlin. J. Geophys. Res.-Atmos. 2001, 106, 8013–8025. [Google Scholar] [CrossRef]
  242. Schindler, M. A QSAR for the prediction of rate constants for the reaction of VOCs with nitrate radicals. Chemosphere 2016, 154, 23–33. [Google Scholar] [CrossRef]
  243. Mogensen, D.; Gierens, R.; Crowley, J.N.; Keronen, P.; Smolander, S.; Sogachev, A.; Nölscher, A.C.; Zhou, L.; Kulmala, M.; Tang, M.J.; et al. Simulations of atmospheric OH, O3 and NO3 reactivities within and above the boreal forest. Atmos. Chem. Phys. 2015, 15, 3909–3932. [Google Scholar] [CrossRef]
  244. Xu, L.; Tsona, N.T.; You, B.; Zhang, Y.; Wang, S.; Yang, Z.; Xue, L.; Du, L. NOx enhances secondary organic aerosol formation from nighttime γ-terpinene ozonolysis. Atmos. Environ. 2020, 225, 117375. [Google Scholar] [CrossRef]
  245. Waring, M.S.; Wells, J.R. Volatile organic compound conversion by ozone, hydroxyl radicals, and nitrate radicals in residential indoor air: Magnitudes and impacts of oxidant sources. Atmos. Environ. 2015, 106, 382–391. [Google Scholar] [CrossRef]
  246. Liu, Y.; Qiu, P.; Li, C.; Li, X.; Ma, W.; Yin, S.; Yu, Q.; Li, J.; Liu, X. Evolution and variations of atmospheric VOCs and O3 photochemistry during a summer O3 event in a county-level city, Southern China. Atmos. Environ. 2022, 272, 118942. [Google Scholar] [CrossRef]
  247. Sun, J.; Yu, X.; Ling, Z.; Fang, G.; Ming, L.; Zhao, J.; Zou, S.; Guan, H.; Wang, H.; Wang, X.; et al. Roles of photochemical consumption of VOCs on regional background O3 concentration and atmospheric reactivity over the pearl river estuary, Southern China. Sci. Total Environ. 2024, 928, 172321. [Google Scholar] [CrossRef]
  248. Ng, N.L.; Canagaratna, M.R.; Zhang, Q.; Jimenez, J.L.; Tian, J.; Ulbrich, I.M.; Kroll, J.H.; Docherty, K.S.; Chhabra, P.S.; Bahreini, R.; et al. Organic aerosol components observed in Northern Hemispheric datasets from Aerosol Mass Spectrometry. Atmos. Chem. Phys. 2010, 10, 4625–4641. [Google Scholar] [CrossRef]
  249. Atkinson, R.; Arey, J. Atmospheric Degradation of Volatile Organic Compounds. Chem. Rev. 2003, 103, 4605–4638. [Google Scholar] [CrossRef] [PubMed]
  250. Docherty, K.S.; Wu, W.; Lim, Y.B.; Ziemann, P.J. Contributions of Organic Peroxides to Secondary Aerosol Formed from Reactions of Monoterpenes with O3. Environ. Sci. Technol. 2005, 39, 4049–4059. [Google Scholar] [CrossRef]
  251. Kroll, J.H.; Seinfeld, J.H. Chemistry of secondary organic aerosol: Formation and evolution of low-volatility organics in the atmosphere. Atmos. Environ. 2008, 42, 3593–3624. [Google Scholar] [CrossRef]
  252. Orlando, J.J.; Tyndall, G.S. Laboratory studies of organic peroxy radical chemistry: An overview with emphasis on recent issues of atmospheric significance. Chem. Soc. Rev. 2012, 41, 6294–6317. [Google Scholar] [CrossRef]
  253. Berndt, T.; Richters, S.; Kaethner, R.; Voigtländer, J.; Stratmann, F.; Sipilä, M.; Kulmala, M.; Herrmann, H. Gas-Phase Ozonolysis of Cycloalkenes: Formation of Highly Oxidized RO2 Radicals and Their Reactions with NO, NO2, SO2, and Other RO2 Radicals. J. Phys. Chem. A 2015, 119, 10336–10348. [Google Scholar] [CrossRef] [PubMed]
  254. Calfapietra, C.; Fares, S.; Manes, F.; Morani, A.; Sgrigna, G.; Loreto, F. Role of Biogenic Volatile Organic Compounds (BVOC) emitted by urban trees on ozone concentration in cities: A review. Environ. Pollut. 2013, 183, 71–80. [Google Scholar] [CrossRef]
  255. Chen, W.; Guenther, A.B.; Jia, S.; Mao, J.; Yan, F.; Wang, X.; Shao, M. Synergistic effects of biogenic volatile organic compounds and soil nitric oxide emissions on summertime ozone formation in China. Sci. Total Environ. 2022, 828, 154218. [Google Scholar] [CrossRef]
  256. Xie, X.; Shao, M.; Liu, Y.; Lu, S.; Chang, C.-C.; Chen, Z.-M. Estimate of initial isoprene contribution to ozone formation potential in Beijing, China. Atmos. Environ. 2008, 42, 6000–6010. [Google Scholar] [CrossRef]
  257. Kim, S.; Lee, M.; Kim, S.; Choi, S.; Seok, S.; Kim, S. Photochemical characteristics of high and low ozone episodes observed in the Taehwa Forest observatory (TFO) in June 2011 near Seoul South Korea. Asia-Pac. J. Atmos. Sci. 2013, 49, 325–331. [Google Scholar] [CrossRef]
  258. Lin, C.; Feng, X.; Heal, M.R. Temporal persistence of intra-urban spatial contrasts in ambient NO2, O3 and Ox in Edinburgh, UK. Atmos. Pollut. Res. 2016, 7, 734–741. [Google Scholar] [CrossRef]
  259. Pancholi, P.; Kumar, A.; Bikundia, D.S.; Chourasiya, S. An observation of seasonal and diurnal behavior of O3–NOx relationships and local/regional oxidant (OX = O3 + NO2) levels at a semi-arid urban site of western India. Sustain. Environ. Res. 2018, 28, 79–89. [Google Scholar] [CrossRef]
  260. Simon, H.; Fallmann, J.; Kropp, T.; Tost, H.; Bruse, M. Urban Trees and Their Impact on Local Ozone Concentration—A Microclimate Modeling Study. Atmosphere 2019, 10, 154. [Google Scholar] [CrossRef]
  261. Zhang, Y.; Dai, J.; Li, Q.; Chen, T.; Mu, J.; Brasseur, G.; Wang, T.; Xue, L. Biogenic volatile organic compounds enhance ozone production and complicate control efforts: Insights from long-term observations in Hong Kong. Atmos. Environ. 2023, 309, 119917. [Google Scholar] [CrossRef]
  262. Liu, C.; Shi, K. A review on methodology in O3-NOx-VOC sensitivity study. Environ. Pollut. 2021, 291, 118249. [Google Scholar] [CrossRef] [PubMed]
  263. Ren, J.; Guo, F.; Xie, S. Diagnosing ozone–NOx–VOC sensitivity and revealing causes of ozone increases in China based on 2013–2021 satellite retrievals. Atmos. Chem. Phys. 2022, 22, 15035–15047. [Google Scholar] [CrossRef]
  264. Iyer, S.; Reiman, H.; Møller, K.H.; Rissanen, M.P.; Kjaergaard, H.G.; Kurtén, T. Computational Investigation of RO2 + HO2 and RO2 + RO2 Reactions of Monoterpene Derived First-Generation Peroxy Radicals Leading to Radical Recycling. J. Phys. Chem. A 2018, 122, 9542–9552. [Google Scholar] [CrossRef]
  265. Bianchi, F.; Junninen, H.; Bigi, A.; Sinclair, V.A.; Dada, L.; Hoyle, C.R.; Zha, Q.; Yao, L.; Ahonen, L.R.; Bonasoni, P.; et al. Biogenic particles formed in the Himalaya as an important source of free tropospheric aerosols. Nat. Geosci. 2021, 14, 4–9. [Google Scholar] [CrossRef]
  266. Friedman, B.; Farmer, D.K. SOA and gas phase organic acid yields from the sequential photooxidation of seven monoterpenes. Atmos. Environ. 2018, 187, 335–345. [Google Scholar] [CrossRef]
  267. Kulmala, M.; Nieminen, T.; Chellapermal, R.; Makkonen, R.; Bäck, J.; Kerminen, V.-M. Climate Feedbacks Linking the Increasing Atmospheric CO2 Concentration, BVOC Emissions, Aerosols and Clouds in Forest Ecosystems. In Biology, Controls and Models of Tree Volatile Organic Compound Emissions; Niinemets, Ü., Monson, R.K., Eds.; Springer: Dordrecht, The Netherlands, 2013; pp. 489–508. ISBN 978-94-007-6606-8. [Google Scholar]
  268. Jimenez, J.L.; Canagaratna, M.R.; Donahue, N.M.; Prevot, A.S.H.; Zhang, Q.; Kroll, J.H.; DeCarlo, P.F.; Allan, J.D.; Coe, H.; Ng, N.L.; et al. Evolution of Organic Aerosols in the Atmosphere. Science 2009, 326, 1525–1529. [Google Scholar] [CrossRef]
  269. Robinson, A.L.; Donahue, N.M.; Shrivastava, M.K.; Weitkamp, E.A.; Sage, A.M.; Grieshop, A.P.; Lane, T.E.; Pierce, J.R.; Pandis, S.N. Rethinking Organic Aerosols: Semivolatile Emissions and Photochemical Aging. Science 2007, 315, 1259–1262. [Google Scholar] [CrossRef]
  270. Carlton, A.G.; Wiedinmyer, C.; Kroll, J.H. A review of Secondary Organic Aerosol (SOA) formation from isoprene. Atmos. Chem. Phys. 2009, 9, 4987–5005. [Google Scholar] [CrossRef]
  271. Dada, L.; Stolzenburg, D.; Simon, M.; Fischer, L.; Heinritzi, M.; Wang, M.; Xiao, M.; Vogel, A.L.; Ahonen, L.; Amorim, A.; et al. Role of sesquiterpenes in biogenic new particle formation. Sci. Adv. 2023, 9, eadi5297. [Google Scholar] [CrossRef] [PubMed]
  272. Bates, K.H.; Burke, G.J.P.; Cope, J.D.; Nguyen, T.B. Secondary organic aerosol and organic nitrogen yields from the nitrate radical (NO3) oxidation of alpha-pinene from various RO2 fates. Atmos. Chem. Phys. 2022, 22, 1467–1482. [Google Scholar] [CrossRef]
  273. Ziemann, P.J.; Atkinson, R. Kinetics, products, and mechanisms of secondary organic aerosol formation. Chem. Soc. Rev. 2012, 41, 6582–6605. [Google Scholar] [CrossRef] [PubMed]
  274. Ehn, M.; Thornton, J.A.; Kleist, E.; Sipilä, M.; Junninen, H.; Pullinen, I.; Springer, M.; Rubach, F.; Tillmann, R.; Lee, B.; et al. A large source of low-volatility secondary organic aerosol. Nature 2014, 506, 476–479. [Google Scholar] [CrossRef]
  275. Perring, A.E.; Pusede, S.E.; Cohen, R.C. An Observational Perspective on the Atmospheric Impacts of Alkyl and Multifunctional Nitrates on Ozone and Secondary Organic Aerosol. Chem. Rev. 2013, 113, 5848–5870. [Google Scholar] [CrossRef]
  276. Kirkby, J.; Duplissy, J.; Sengupta, K.; Frege, C.; Gordon, H.; Williamson, C.; Heinritzi, M.; Simon, M.; Yan, C.; Almeida, J.; et al. Ion-induced nucleation of pure biogenic particles. Nature 2016, 533, 521–526. [Google Scholar] [CrossRef]
  277. Tröstl, J.; Chuang, W.K.; Gordon, H.; Heinritzi, M.; Yan, C.; Molteni, U.; Ahlm, L.; Frege, C.; Bianchi, F.; Wagner, R.; et al. The role of low-volatility organic compounds in initial particle growth in the atmosphere. Nature 2016, 533, 527–531. [Google Scholar] [CrossRef]
  278. Cash, J.M.; Heal, M.R.; Langford, B.; Drewer, J. A review of stereochemical implications in the generation of secondary organic aerosol from isoprene oxidation. Environ. Sci. Process. Impacts 2016, 18, 1369–1380. [Google Scholar] [CrossRef]
  279. Yang, Z.; Du, L.; Li, Y.; Ge, X. Secondary organic aerosol formation from monocyclic aromatic hydrocarbons: Insights from laboratory studies. Environ. Sci. Process. Impacts 2022, 24, 351–379. [Google Scholar] [CrossRef]
  280. Gao, Y.; Yan, F.; Ma, M.; Ding, A.; Liao, H.; Wang, S.; Wang, X.; Zhao, B.; Cai, W.; Su, H.; et al. Unveiling the dipole synergic effect of biogenic and anthropogenic emissions on ozone concentrations. Sci. Total Environ. 2022, 818, 151722. [Google Scholar] [CrossRef]
  281. Shrivastava, M.; Cappa, C.D.; Fan, J.; Goldstein, A.H.; Guenther, A.B.; Jimenez, J.L.; Kuang, C.; Laskin, A.; Martin, S.T.; Ng, N.L.; et al. Recent advances in understanding secondary organic aerosol: Implications for global climate forcing. Rev. Geophys. 2017, 55, 509–559. [Google Scholar] [CrossRef]
  282. Fierravanti, A.; Fierravanti, E.; Cocozza, C.; Tognetti, R.; Rossi, S. Eligible reference cities in relation to BVOC-derived O3 pollution. Urban For. Urban Green. 2017, 28, 73–80. [Google Scholar] [CrossRef]
  283. Yenisoy-Karakaş, S.; Dörter, M.; Odabasi, M. Intraday and interday variations of 69 volatile organic compounds (BVOCs and AVOCs) and their source profiles at a semi-urban site. Sci. Total Environ. 2020, 723, 138028. [Google Scholar] [CrossRef] [PubMed]
  284. Li, N.; He, Q.; Greenberg, J.; Guenther, A.; Li, J.; Cao, J.; Wang, J.; Liao, H.; Wang, Q.; Zhang, Q. Impacts of biogenic and anthropogenic emissions on summertime ozone formation in the Guanzhong Basin, China. Atmos. Chem. Phys. 2018, 18, 7489–7507. [Google Scholar] [CrossRef]
  285. Choi, J.-S.; Choi, Y.; Ahn, J.; Park, J.; Oh, J.; Lee, G.; Park, T.; Park, G.; Owen, J.S.; Lee, T. Observation of Secondary Organic Aerosol and New Particle Formation at a Remote Site in Baengnyeong Island, Korea. Asian J. Atmos. Environ. 2017, 11, 300–312. [Google Scholar] [CrossRef]
  286. Ghirardo, A.; Xie, J.; Zheng, X.; Wang, Y.; Grote, R.; Block, K.; Wildt, J.; Mentel, T.; Kiendler-Scharr, A.; Hallquist, M.; et al. Urban stress-induced biogenic VOC emissions and SOA-forming potentials in Beijing. Atmos. Chem. Phys. 2016, 16, 2901–2920. [Google Scholar] [CrossRef]
  287. Šimpraga, M.; Ghimire, R.P.; Van Der Straeten, D.; Blande, J.D.; Kasurinen, A.; Sorvari, J.; Holopainen, T.; Adriaenssens, S.; Holopainen, J.K.; Kivimäenpää, M. Unravelling the functions of biogenic volatiles in boreal and temperate forest ecosystems. Eur. J. Forest Res. 2019, 138, 763–787. [Google Scholar] [CrossRef]
  288. Yong, J.; Xie, Y.; Guo, H.; Li, Y.; Sun, S. Unraveling the influence of biogenic volatile organic compounds and their constituents on ozone and SOA formation within the Yellow River Basin, China. Chemosphere 2024, 353, 141549. [Google Scholar] [CrossRef]
  289. Jia, Y.; Qiao, L.; Xie, W.; Li, L. Spatial-temporal characteristics of ambient isoprene and monoterpene and their ozone and secondary organic aerosol formation potential in China. Environ. Rev. 2024, 32, 203–213. [Google Scholar] [CrossRef]
  290. Zheng, J.; Shao, M.; Che, W.; Zhang, L.; Zhong, L.; Zhang, Y.; Streets, D. Speciated VOC Emission Inventory and Spatial Patterns of Ozone Formation Potential in the Pearl River Delta, China. Environ. Sci. Technol. 2009, 43, 8580–8586. [Google Scholar] [CrossRef]
  291. Ryu, Y.-H.; Baik, J.-J.; Kwak, K.-H.; Kim, S.; Moon, N. Impacts of urban land-surface forcing on ozone air quality in the Seoul metropolitan area. Atmos. Chem. Phys. 2013, 13, 2177–2194. [Google Scholar] [CrossRef]
  292. Li, Y.; Yin, S.; Yu, S.; Bai, L.; Wang, X.; Lu, X.; Ma, S. Characteristics of ozone pollution and the sensitivity to precursors during early summer in central plain, China. J. Environ. Sci. 2021, 99, 354–368. [Google Scholar] [CrossRef] [PubMed]
  293. Mahilang, M.; Deb, M.K.; Pervez, S. Biogenic secondary organic aerosols: A review on formation mechanism, analytical challenges and environmental impacts. Chemosphere 2021, 262, 127771. [Google Scholar] [CrossRef]
  294. Yan, F.; Chen, W.; Jia, S.; Zhong, B.; Yang, L.; Mao, J.; Chang, M.; Shao, M.; Yuan, B.; Situ, S.; et al. Stabilization for the secondary species contribution to PM2.5 in the Pearl River Delta (PRD) over the past decade, China: A meta-analysis. Atmos. Environ. 2020, 242, 117817. [Google Scholar] [CrossRef]
  295. Déméautis, T.; Delles, M.; Tomaz, S.; Monneret, G.; Glehen, O.; Devouassoux, G.; George, C.; Bentaher, A. Pathogenic Mechanisms of Secondary Organic Aerosols. Chem. Res. Toxicol. 2022, 35, 1146–1161. [Google Scholar] [CrossRef] [PubMed]
  296. Zeng, J.; Zhang, Y.; Zhang, H.; Song, W.; Wu, Z.; Wang, X. Design and characterization of a semi-open dynamic chamber for measuring biogenic volatile organic compound (BVOC) emissions from plants. Atmos. Meas. Tech. 2022, 15, 79–93. [Google Scholar] [CrossRef]
  297. Xu, Y.; Ling, L.; Xie, P.; Hu, R.; Li, Z.; Chen, H. Low cost and high resolution electrochemical sensing for atmospheric NO2 measurement. Huanjing Kexue Xuebao/Acta Sci. Circumstantiae 2018, 38, 3488–3495. [Google Scholar] [CrossRef]
  298. Kim, S.; Guenther, A.; Apel, E. Quantitative and qualitative sensing techniques for biogenic volatile organic compounds and their oxidation products. Environ. Sci. Process. Impacts 2013, 15, 1301–1314. [Google Scholar] [CrossRef]
Figure 1. Narrative review flow diagram.
Figure 1. Narrative review flow diagram.
Toxics 13 00364 g001
Figure 2. Distribution of 507 studies on biogenic volatile organic compounds (BVOCs) since 1999. (a) The change in the number of publications related to BVOCs in the world and the proportion of the number of articles published in China and other countries. (b) The quantitative distribution of relevant publications based on research areas.
Figure 2. Distribution of 507 studies on biogenic volatile organic compounds (BVOCs) since 1999. (a) The change in the number of publications related to BVOCs in the world and the proportion of the number of articles published in China and other countries. (b) The quantitative distribution of relevant publications based on research areas.
Toxics 13 00364 g002
Figure 3. The characteristics of different model simulation methods of biogenic volatile organic compounds (BVOCs) and their applicable scenarios.
Figure 3. The characteristics of different model simulation methods of biogenic volatile organic compounds (BVOCs) and their applicable scenarios.
Toxics 13 00364 g003
Figure 4. The effects of plant internal and environmental factors on biogenic volatile organic compounds (BVOCs)(notes: MTS, monoterpenes; ISO, isoprene; OBVOCs, oxygenated biogenic volatile organic compounds; MEP, Methylerythritol Phosphate Pathway; MVA, Mevalonate Pathway; LOX, Lipoxygenase Pathway; AVOCs, anthropogenic volatile organic compounds; CO2, carbon dioxide; O3, ozone).
Figure 4. The effects of plant internal and environmental factors on biogenic volatile organic compounds (BVOCs)(notes: MTS, monoterpenes; ISO, isoprene; OBVOCs, oxygenated biogenic volatile organic compounds; MEP, Methylerythritol Phosphate Pathway; MVA, Mevalonate Pathway; LOX, Lipoxygenase Pathway; AVOCs, anthropogenic volatile organic compounds; CO2, carbon dioxide; O3, ozone).
Toxics 13 00364 g004
Figure 5. The initial oxidation reaction of biogenic volatile organic compounds (BVOCs) and the pathway diagram of their formation of ozone (O3) and secondary organic aerosols (SOAs) (notes: for the corresponding abbreviations in the chart, please refer to nomenclature (Table S1)).
Figure 5. The initial oxidation reaction of biogenic volatile organic compounds (BVOCs) and the pathway diagram of their formation of ozone (O3) and secondary organic aerosols (SOAs) (notes: for the corresponding abbreviations in the chart, please refer to nomenclature (Table S1)).
Toxics 13 00364 g005
Table 1. Comprehensive comparisons of biogenic volatile organic compounds (BVOCs) emission measurement methods.
Table 1. Comprehensive comparisons of biogenic volatile organic compounds (BVOCs) emission measurement methods.
Measurement MethodsOn-LineOff-Line
Main technologiesDirect analysis:
PTR-MS [32]
FTIR [33]
GC-FID [34]
Portable GC/MS [35]
E-noses [36]
Sampling:
sampling bags,
adsorption tubes,
sampling tanks
Analysis:
GC-MS [37]
TDS-GC/MS [38]
TCT-GC/MS [39]
TD-GC/TOF-MS [40]
SPME-GC/MS [41]
AdvantagesReal-time monitoring;
high temporal resolution;
high sensitivity;
high degree of automation
Lower equipment costs;
high flexibility;
diverse and detailed analysis
DisadvantagesExpensive equipment;
complexity;
high on-site requirements
Time delay;
lower temporal resolution;
human error
Comprehensive comparison
  • Precision and sensitivity: on-line > off-line
  • Operational complexity: on-line > off-line
Table 2. Biogenic volatile organic compounds (BVOCs) analytical techniques comparison (notes: Ref., references).
Table 2. Biogenic volatile organic compounds (BVOCs) analytical techniques comparison (notes: Ref., references).
Analysis TechniqueDetector TypeSensitivityTime
Resolution
Advantages and DisadvantagesApplicable BVOC TypesRef.
PTR-MSMass spectrometryppt- ppmReal-time analysisSensitivity and costMost BVOCs [42,43]
FTIRPyroelectric detector or photoconductive detectorppb- ppmMinutes to hoursVersatility and sensitivityPolar BVOCs [44,45]
GC-FIDFlame ionization detectionpptMinutesHigh speed and limitationsNonpolar BVOCs [34,46]
E-nosesGas sensorsppb- ppmMinutesSpeed and resolutionTrace BVOCs [36,47]
Portable GC/MSMass spectrometerppbMinutesPortability and limitationsMost BVOCs [48,49]
GC-MSMass spectrometerpptMinutes to hoursPrecision and complexityMost BVOCs [37,50]
TD-GC/MSMass spectrometerppbMinutesSensitivity and time-consumingTrace BVOCs [38,51]
TCT-GC/MSMass spectrometerppbMinutesPrecision and requirementsComplex BVOCs [39,52]
TD-GC/TOF-MSTime-of-flight mass spectrometerppbMinutesResolution and costTrace BVOCs [40,53]
SPME-
GC/MS
Mass spectrometerppbMinutesSimplicity and sensitivityMost BVOCs [41,54]
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

Luo, R.; Lun, X.; Gao, R.; Wang, L.; Yang, Y.; Su, X.; Habibullah-Al-Mamun, M.; Xu, X.; Li, H.; Li, J. A Review of Biogenic Volatile Organic Compounds from Plants: Research Progress and Future Prospects. Toxics 2025, 13, 364. https://doi.org/10.3390/toxics13050364

AMA Style

Luo R, Lun X, Gao R, Wang L, Yang Y, Su X, Habibullah-Al-Mamun M, Xu X, Li H, Li J. A Review of Biogenic Volatile Organic Compounds from Plants: Research Progress and Future Prospects. Toxics. 2025; 13(5):364. https://doi.org/10.3390/toxics13050364

Chicago/Turabian Style

Luo, Rongrong, Xiaoxiu Lun, Rui Gao, Le Wang, Yuan Yang, Xingqian Su, Md Habibullah-Al-Mamun, Xiaohang Xu, Hong Li, and Jinjuan Li. 2025. "A Review of Biogenic Volatile Organic Compounds from Plants: Research Progress and Future Prospects" Toxics 13, no. 5: 364. https://doi.org/10.3390/toxics13050364

APA Style

Luo, R., Lun, X., Gao, R., Wang, L., Yang, Y., Su, X., Habibullah-Al-Mamun, M., Xu, X., Li, H., & Li, J. (2025). A Review of Biogenic Volatile Organic Compounds from Plants: Research Progress and Future Prospects. Toxics, 13(5), 364. https://doi.org/10.3390/toxics13050364

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