Volatile organic compound

Last updated

Volatile organic compounds (VOCs) are organic compounds that have a high vapor pressure at room temperature. [1] High vapor pressure correlates with a low boiling point, which relates to the number of the sample's molecules in the surrounding air, a trait known as volatility. [2]

Contents

VOCs are responsible for the odor of scents and perfumes as well as pollutants. VOCs play an important role in communication between animals and plants, e.g. attractants for pollinators, [3] protection from predation, [4] and even inter-plant interactions. [5] Some VOCs are dangerous to human health or cause harm to the environment. Anthropogenic VOCs are regulated by law, especially indoors, where concentrations are the highest. Most VOCs are not acutely toxic, but may have long-term chronic health effects. Some VOCs have been used in pharmacy, while others are target of administrative controls because of their recreational use.

Definitions

Diverse definitions of the term VOC are in use. Some examples are presented below.

Canada

Health Canada classifies VOCs as organic compounds that have boiling points roughly in the range of 50 to 250 °C (122 to 482 °F). The emphasis is placed on commonly encountered VOCs that would have an effect on air quality. [6]

European Union

The European Union defines a VOC as "any organic compound as well as the fraction of creosote, having at 293.15 K a vapour pressure of 0,01 kPa or more, or having a corresponding volatility under the particular conditions of use;" [7] . The VOC Solvents Emissions Directive was the main policy instrument for the reduction of industrial emissions of volatile organic compounds (VOCs) in the European Union. It covers a wide range of solvent-using activities, e.g. printing, surface cleaning, vehicle coating, dry cleaning and manufacture of footwear and pharmaceutical products. The VOC Solvents Emissions Directive requires installations in which such activities are applied to comply either with the emission limit values set out in the Directive or with the requirements of the so-called reduction scheme. Article 13 of The Paints Directive, approved in 2004, amended the original VOC Solvents Emissions Directive and limits the use of organic solvents in decorative paints and varnishes and in vehicle finishing products. The Paints Directive sets out maximum VOC content limit values for paints and varnishes in certain applications. [8] [9] The Solvents Emissions Directive was replaced by the Industrial Emissions Directive from 2013.

China

The People's Republic of China defines a VOC as those compounds that have "originated from automobiles, industrial production and civilian use, burning of all types of fuels, storage and transportation of oils, fitment finish, coating for furniture and machines, cooking oil fume and fine particles (PM 2.5)", and similar sources. [10] The Three-Year Action Plan for Winning the Blue Sky Defence War released by the State Council in July 2018 creates an action plan to reduce 2015 VOC emissions 10% by 2020. [11]

India

The Central Pollution Control Board of India released the Air (Prevention and Control of Pollution) Act in 1981, amended in 1987, to address concerns about air pollution in India. [12] While the document does not differentiate between VOCs and other air pollutants, the CPCB monitors "oxides of nitrogen (NOx), sulphur dioxide (SO2), fine particulate matter (PM10) and suspended particulate matter (SPM)". [13]

United States

Thermal oxidizers provide an air pollution abatement option for VOCs from industrial air flows. A thermal oxidizer is an EPA-approved device to treat VOCs. Thermal-oxidizer-rto.jpg
Thermal oxidizers provide an air pollution abatement option for VOCs from industrial air flows. A thermal oxidizer is an EPA-approved device to treat VOCs.

The definitions of VOCs used for control of precursors of photochemical smog used by the U.S. Environmental Protection Agency (EPA) and state agencies in the US with independent outdoor air pollution regulations include exemptions for VOCs that are determined to be non-reactive, or of low-reactivity in the smog formation process. Prominent is the VOC regulation issued by the South Coast Air Quality Management District in California and by the California Air Resources Board (CARB). [15] However, this specific use of the term VOCs can be misleading, especially when applied to indoor air quality because many chemicals that are not regulated as outdoor air pollution can still be important for indoor air pollution.

Following a public hearing in September 1995, California's ARB uses the term "reactive organic gases" (ROG) to measure organic gases. The CARB revised the definition of "Volatile Organic Compounds" used in their consumer products regulations, based on the committee's findings. [16]

In addition to drinking water, VOCs are regulated in pollutant discharges to surface waters (both directly and via sewage treatment plants) [17] as hazardous waste, [18] but not in non-industrial indoor air. [19] The Occupational Safety and Health Administration (OSHA) regulates VOC exposure in the workplace. Volatile organic compounds that are classified as hazardous materials are regulated by the Pipeline and Hazardous Materials Safety Administration while being transported.

Biologically generated VOCs

Most VOCs in Earth's atmosphere are biogenic, largely emitted by plants. [2]

Major biogenic VOCs [20]
compoundrelative contributionamount emitted (Tg/y)
isoprene 62.2%594±34
terpenes 10.9%95±3
pinene isomers5.6%48.7±0.8
sesquiterpenes 2.4%20±1
methanol 6.4%130±4

Biogenic volatile organic compounds (BVOCs) encompass VOCs emitted by plants, animals, or microorganisms, and while extremely diverse, are most commonly terpenoids, alcohols, and carbonyls (methane and carbon monoxide are generally not considered). [21] Not counting methane, biological sources emit an estimated 760 teragrams of carbon per year in the form of VOCs. [20] The majority of VOCs are produced by plants, the main compound being isoprene. Small amounts of VOCs are produced by animals and microbes. [22] Many VOCs are considered secondary metabolites, which often help organisms in defense, such as plant defense against herbivory. The strong odor emitted by many plants consists of green leaf volatiles, a subset of VOCs. Although intended for nearby organisms to detect and respond to, these volatiles can be detected and communicated through wireless electronic transmission, by embedding nanosensors and infrared transmitters into the plant materials themselves. [23]

Emissions are affected by a variety of factors, such as temperature, which determines rates of volatilization and growth, and sunlight, which determines rates of biosynthesis. Emission occurs almost exclusively from the leaves, the stomata in particular. VOCs emitted by terrestrial forests are often oxidized by hydroxyl radicals in the atmosphere; in the absence of NOx pollutants, VOC photochemistry recycles hydroxyl radicals to create a sustainable biosphere-atmosphere balance. [24] Due to recent climate change developments, such as warming and greater UV radiation, BVOC emissions from plants are generally predicted to increase, thus upsetting the biosphere-atmosphere interaction and damaging major ecosystems. [25] A major class of VOCs is the terpene class of compounds, such as myrcene. [26]

Providing a sense of scale, a forest 62,000 square kilometres (24,000 sq mi) in area, the size of the US state of Pennsylvania, is estimated to emit 3,400,000 kilograms (7,500,000 lb) of terpenes on a typical August day during the growing season. [27] Researchers investigating mechanisms of induction of genes producing volatile organic compounds, and the subsequent increase in volatile terpenes, has been achieved in maize using (Z)-3-hexen-1-ol and other plant hormones. [28]

Anthropogenic sources

The handling of petroleum-based fuels is a major source of VOCs. Fuel dispenser in use.jpg
The handling of petroleum-based fuels is a major source of VOCs.

Anthropogenic sources emit about 142 teragrams (1.42 × 1011 kg) of carbon per year in the form of VOCs. [29]

The major source of man-made VOCs are: [30]

Indoor VOCs

Concentrations of VOCs in indoor air may be 2 to 5 times greater than in outdoor air, sometimes far greater. [19] During certain activities, indoor levels of VOCs may reach 1,000 times that of the outside air. Studies have shown that emissions of individual VOC species are not that high in an indoor environment, but the total concentration of all VOCs (TVOC) indoors can be up to five times higher than that of outdoor levels. [33]

New buildings experience particularly high levels of VOC off-gassing indoors because of the abundant new materials (building materials, fittings, surface coverings and treatments such as glues, paints and sealants) exposed to the indoor air, emitting multiple VOC gases. [34] This off-gassing has a multi-exponential decay trend that is discernible over at least two years, with the most volatile compounds decaying with a time-constant of a few days, and the least volatile compounds decaying with a time-constant of a few years. [35]

New buildings may require intensive ventilation for the first few months, or a bake-out treatment. Existing buildings may be replenished with new VOC sources, such as new furniture, consumer products, and redecoration of indoor surfaces, all of which lead to a continuous background emission of TVOCs, and requiring improved ventilation. [34]

Numerous studies [35] show strong seasonal variations in indoors VOC emissions, with emission rates increasing in summer. This is largely due to the rate of diffusion of VOC species through materials to the surface, increasing with temperature. Most studies have shown that this leads to generally higher concentrations of TVOCs indoors in summer. [35]

Indoor air quality measurements

Measurement of VOCs from the indoor air is done with sorption tubes e. g. Tenax (for VOCs and SVOCs) or DNPH-cartridges (for carbonyl-compounds) or air detector. The VOCs adsorb on these materials and are afterwards desorbed either thermally (Tenax) or by elution (DNPH) and then analyzed by GC-MS/FID or HPLC. Reference gas mixtures are required for quality control of these VOC-measurements. [36] Furthermore, VOC emitting products used indoors, e.g. building products and furniture, are investigated in emission test chambers under controlled climatic conditions. [37] For quality control of these measurements round robin tests are carried out, therefore reproducibly emitting reference materials are ideally required. [36] Other methods have used proprietary Silcosteel-coated canisters with constant flow inlets to collect samples over several days. [38] These methods are not limited by the adsorbing properties of materials like Tenax.

Regulation of indoor VOC emissions

In most countries, a separate definition of VOCs is used with regard to indoor air quality that comprises each organic chemical compound that can be measured as follows: adsorption from air on Tenax TA, thermal desorption, gas chromatographic separation over a 100% nonpolar column (dimethylpolysiloxane). VOC (volatile organic compounds) are all compounds that appear in the gas chromatogram between and including n-hexane and n-hexadecane. Compounds appearing earlier are called VVOC (very volatile organic compounds); compounds appearing later are called SVOC (semi-volatile organic compounds).

France, Germany (AgBB/DIBt), Belgium, Norway (TEK regulation), and Italy (CAM Edilizia) have enacted regulations to limit VOC emissions from commercial products. European industry has developed numerous voluntary ecolabels and rating systems, such as EMICODE, [39] M1, [40] Blue Angel, [41] GuT (textile floor coverings), [42] Nordic Swan Ecolabel, [43] EU Ecolabel, [44] and Indoor Air Comfort. [45] In the United States, several standards exist; California Standard CDPH Section 01350 [46] is the most common one. These regulations and standards changed the marketplace, leading to an increasing number of low-emitting products.

Health risks

Respiratory, allergic, or immune effects in infants or children are associated with man-made VOCs and other indoor or outdoor air pollutants. [47]

Some VOCs, such as styrene and limonene, can react with nitrogen oxides or with ozone to produce new oxidation products and secondary aerosols, which can cause sensory irritation symptoms. [48] VOCs contribute to the formation of tropospheric ozone and smog. [49] [50]

Health effects include eye, nose, and throat irritation; headaches, loss of coordination, nausea; and damage to the liver, kidney, and central nervous system. [51] Some organics can cause cancer in animals; some are suspected or known to cause cancer in humans. Key signs or symptoms associated with exposure to VOCs include conjunctival irritation, nose and throat discomfort, headache, allergic skin reaction, dyspnea, declines in serum cholinesterase levels, nausea, vomiting, nose bleeding, fatigue, dizziness. [52]

The ability of organic chemicals to cause health effects varies greatly from those that are highly toxic to those with no known health effects. As with other pollutants, the extent and nature of the health effect will depend on many factors including level of exposure and length of time exposed. Eye and respiratory tract irritation, headaches, dizziness, visual disorders, and memory impairment are among the immediate symptoms that some people have experienced soon after exposure to some organics. At present, not much is known about what health effects occur from the levels of organics usually found in homes. [53]

Ingestion

While null in comparison to the concentrations found in indoor air, benzene, toluene, and methyl tert-butyl ether (MTBE) were found in samples of human milk and increase the concentrations of VOCs that we are exposed to throughout the day. [54] A study notes the difference between VOCs in alveolar breath and inspired air suggesting that VOCs are ingested, metabolized, and excreted via the extra-pulmonary pathway. [55] VOCs are also ingested by drinking water in varying concentrations. Some VOC concentrations were over the EPA's National Primary Drinking Water Regulations and China's National Drinking Water Standards set by the Ministry of Ecology and Environment. [56]

Dermal absorption

The presence of VOCs in the air and in groundwater has prompted more studies. Several studies have been performed to measure the effects of dermal absorption of specific VOCs. Dermal exposure to VOCs like formaldehyde and toluene downregulate antimicrobial peptides on the skin like cathelicidin LL-37, human β-defensin 2 and 3. [57]  Xylene and formaldehyde worsen allergic inflammation in animal models. [58] Toluene also increases the dysregulation of filaggrin: a key protein in dermal regulation. [59] this was confirmed by immunofluorescence to confirm protein loss and western blotting to confirm mRNA loss. These experiments were done on human skin samples. Toluene exposure also decreased the water in the trans-epidermal layer allowing for vulnerability in the skin's layers. [57] [60]

Limit values for VOC emissions

Limit values for VOC emissions into indoor air are published by AgBB, [61] AFSSET, California Department of Public Health, and others. These regulations have prompted several companies in the paint and adhesive industries to adapt with VOC level reductions their products.[ citation needed ] VOC labels and certification programs may not properly assess all of the VOCs emitted from the product, including some chemical compounds that may be relevant for indoor air quality. [62] Each ounce of colorant added to tint paint may contain between 5 and 20 grams of VOCs. A dark color, however, could require 5–15 ounces of colorant, adding up to 300 or more grams of VOCs per gallon of paint. [63]

VOCs in healthcare settings

VOCs are also found in hospital and health care environments. In these settings, these chemicals are widely used for cleaning, disinfection, and hygiene of the different areas. [64] Thus, health professionals such as nurses, doctors, sanitation staff, etc., may present with adverse health effects such as asthma; however, further evaluation is required to determine the exact levels and determinants that influence the exposure to these compounds. [64] [65] [66]

Studies have shown that the concentration levels of different VOCs such as halogenated and aromatic hydrocarbons differ substantially between areas of the same hospital. However, one of these studies reported that ethanol, isopropanol, ether, and acetone were the main compounds in the interior of the site. [67] [68] Following the same line, in a study conducted in the United States, it was established that nursing assistants are the most exposed to compounds such as ethanol, while medical equipment preparers are most exposed to 2-propanol. [67] [68]

In relation to exposure to VOCs by cleaning and hygiene personnel, a study conducted in 4 hospitals in the United States established that sterilization and disinfection workers are linked to exposures to d-limonene and 2-propanol, while those responsible for cleaning with chlorine-containing products are more likely to have higher levels of exposure to α-pinene and chloroform. [66] Those who perform floor and other surface cleaning tasks (e.g., floor waxing) and who use quaternary ammonium, alcohol, and chlorine-based products are associated with a higher VOC exposure than the two previous groups, that is, they are particularly linked to exposure to acetone, chloroform, α-pinene, 2-propanol or d-limonene. [66]

Other healthcare environments such as nursing and age care homes have been rarely a subject of study, even though the elderly and vulnerable populations may spend considerable time in these indoor settings where they might be exposed to VOCs, derived from the common use of cleaning agents, sprays and fresheners. [69] [70] In a study conducted in France, a team of researchers developed an online questionnaire for different social and age care facilities, asking about cleaning practices, products used, and the frequency of these activities. As a result, more than 200 chemicals were identified, of which 41 are known to have adverse health effects, 37 of them being VOCs. The health effects include skin sensitization, reproductive and organ-specific toxicity, carcinogenicity, mutagenicity, and endocrine-disrupting properties. [69] Furthermore, in another study carried out in the same European country, it was found that there is a significant association between breathlessness in the elderly population and elevated exposure to VOCs such as toluene and o-xylene, unlike the remainder of the population. [71]

Analytical methods

Sampling

Obtaining samples for analysis is challenging. VOCs, even when at dangerous levels, are dilute, so preconcentration is typically required. Many components of the atmosphere are mutually incompatible, e.g. ozone and organic compounds, peroxyacyl nitrates and many organic compounds. Furthermore, collection of VOCs by condensation in cold traps also accumulates a large amount of water, which generally must be removed selectively, depending on the analytical techniques to be employed. [30] Solid-phase microextraction (SPME) techniques are used to collect VOCs at low concentrations for analysis. [72] As applied to breath analysis, the following modalities are employed for sampling: gas sampling bags, syringes, evacuated steel and glass containers. [73]

Principle and measurement methods

In the U.S., standard methods have been established by the National Institute for Occupational Safety and Health (NIOSH) and another by U.S. OSHA. Each method uses a single component solvent; butanol and hexane cannot be sampled, however, on the same sample matrix using the NIOSH or OSHA method. [74]

VOCs are quantified and identified by two broad techniques. The major technique is gas chromatography (GC). GC instruments allow the separation of gaseous components. When coupled to a flame ionization detector (FID) GCs can detect hydrocarbons at the parts per trillion levels. Using electron capture detectors, GCs are also effective for organohalide such as chlorocarbons.

The second major technique associated with VOC analysis is mass spectrometry, which is usually coupled with GC, giving the hyphenated technique of GC-MS. [75]

Direct injection mass spectrometry techniques are frequently utilized for the rapid detection and accurate quantification of VOCs. [76] PTR-MS is among the methods that have been used most extensively for the on-line analysis of biogenic and anthropogenic VOCs. [77] PTR-MS instruments based on time-of-flight mass spectrometry have been reported to reach detection limits of 20 pptv after 100 ms and 750 ppqv after 1 min. measurement (signal integration) time. The mass resolution of these devices is between 7000 and 10,500 m/Δm, thus it is possible to separate most common isobaric VOCs and quantify them independently. [78]

Chemical fingerprinting and breath analysis

The exhaled human breath contains a few thousand volatile organic compounds and is used in breath biopsy to serve as a VOC biomarker to test for diseases, [73] such as lung cancer. [79] One study has shown that "volatile organic compounds ... are mainly blood borne and therefore enable monitoring of different processes in the body." [80] And it appears that VOC compounds in the body "may be either produced by metabolic processes or inhaled/absorbed from exogenous sources" such as environmental tobacco smoke. [79] [81] Chemical fingerprinting and breath analysis of volatile organic compounds has also been demonstrated with chemical sensor arrays, which utilize pattern recognition for detection of component volatile organics in complex mixtures such as breath gas.

Metrology for VOC measurements

To achieve comparability of VOC measurements, reference standards traceable to SI-units are required. For a number of VOCs gaseous reference standards are available from specialty gas suppliers or national metrology institutes, either in the form of cylinders or dynamic generation methods. However, for many VOCs, such as oxygenated VOCs, monoterpenes, or formaldehyde, no standards are available at the appropriate amount of fraction due to the chemical reactivity or adsorption of these molecules. Currently, several national metrology institutes are working on the lacking standard gas mixtures at trace level concentration, minimising adsorption processes, and improving the zero gas. [36] The final scopes are for the traceability and the long-term stability of the standard gases to be in accordance with the data quality objectives (DQO, maximum uncertainty of 20% in this case) required by the WMO/GAW program. [82]

See also

Related Research Articles

<span class="mw-page-title-main">Smoke</span> Mass of airborne particulates and gases

Smoke is a suspension of airborne particulates and gases emitted when a material undergoes combustion or pyrolysis, together with the quantity of air that is entrained or otherwise mixed into the mass. It is commonly an unwanted by-product of fires, but may also be used for pest control (fumigation), communication, defensive and offensive capabilities in the military, cooking, or smoking. It is used in rituals where incense, sage, or resin is burned to produce a smell for spiritual or magical purposes. It can also be a flavoring agent and preservative.

<span class="mw-page-title-main">Formaldehyde</span> Organic compound (H–CHO); simplest aldehyde

Formaldehyde ( for-MAL-di-hide, fər-) (systematic name methanal) is an organic compound with the chemical formula CH2O and structure H−CHO, more precise H2C=O. The compound is a pungent, colourless gas that polymerises spontaneously into paraformaldehyde. It is stored as aqueous solutions (formalin), which consists mainly of the hydrate CH2(OH)2. It is the simplest of the aldehydes (R−CHO). As a precursor to many other materials and chemical compounds, in 2006 the global production of formaldehyde was estimated at 12 million tons per year. It is mainly used in the production of industrial resins, e.g., for particle board and coatings. Small amounts also occur naturally.

<span class="mw-page-title-main">Indoor air quality</span> Air quality within and around buildings and structures

Indoor air quality (IAQ) is the air quality within and around buildings and structures. IAQ is known to affect the health, comfort, and well-being of building occupants. Poor indoor air quality has been linked to sick building syndrome, reduced productivity, and impaired learning in schools. Common pollutants of indoor air include: secondhand tobacco smoke, air pollutants from indoor combustion, radon, molds and other allergens, carbon monoxide, volatile organic compounds, legionella and other bacteria, asbestos fibers, carbon dioxide, ozone and particulates. Source control, filtration, and the use of ventilation to dilute contaminants are the primary methods for improving indoor air quality.

<span class="mw-page-title-main">Dichloromethane</span> Chemical compound

Dichloromethane is an organochlorine compound with the formula CH2Cl2. This colorless, volatile liquid with a chloroform-like, sweet odor is widely used as a solvent. Although it is not miscible with water, it is slightly polar, and miscible with many organic solvents.

<span class="mw-page-title-main">Exhaust gas</span> Gases emitted as a result of fuel reactions in combustion engines

Exhaust gas or flue gas is emitted as a result of the combustion of fuels such as natural gas, gasoline (petrol), diesel fuel, fuel oil, biodiesel blends, or coal. According to the type of engine, it is discharged into the atmosphere through an exhaust pipe, flue gas stack, or propelling nozzle. It often disperses downwind in a pattern called an exhaust plume.

<span class="mw-page-title-main">Air freshener</span> Product used to mask odors

Air fresheners are products designed to reduce unwanted odors in indoor spaces, or to introduce pleasant fragrances, or both. They typically emit fragrance to mask odors, but may use other methods of action such as absorbing, bonding to, or chemically altering compounds in the air that produce smells, killing organisms that produce smells, or disrupting the sense of smell to reduce perception of unpleasant smells.

<span class="mw-page-title-main">Persistent organic pollutant</span> Organic compounds that are resistant to environmental degradation

Persistent organic pollutants (POPs) are organic compounds that are resistant to degradation through chemical, biological, and photolytic processes. They are toxic chemicals that adversely affect human health and the environment around the world. Because they can be transported by wind and water, most POPs generated in one country can and do affect people and wildlife far from where they are used and released.

Difluoromethane, also called difluoromethylene, HFC-32Methylene Fluoride or R-32, is an organic compound of the dihalogenoalkane variety. It has the formula of CH2F2. It is a colorless gas in the ambient atmosphere and is slightly soluble in the water, with a high thermal stability. Due to the low melting and boiling point, (-136.0 °C and -51.6 °C respectively) contact with this compound may result in frostbite. In the United States, the Clean Air Act Section 111 on Volatile Organic Compounds (VOC) has listed difluoromethane as an exception (since 1997) from the definition of VOC due to its low production of tropospheric ozone. Difluoromethane is commonly used in endothermic processes such as refrigeration or air conditioning.

<span class="mw-page-title-main">Non-methane volatile organic compound</span>

Non-methane volatile organic compounds (NMVOCs) are a set of organic compounds that are typically photochemically reactive in the atmosphere—marked by the exclusion of methane. NMVOCs include a large variety of chemically different compounds, such as benzene, ethanol, formaldehyde, cyclohexane, 1,1,1-trichloroethane and acetone. Essentially, NMVOCs are identical to volatile organic compounds (VOCs), but with methane excluded. Methane is excluded in air-pollution contexts because it is not toxic. It is however a very potent greenhouse gas, with low reactivity and thus a long lifetime in the atmosphere. An important subset of NMVOCs are the non-methane hydrocarbons (NMHCs).

<span class="mw-page-title-main">2-Ethylhexanol</span> Chemical compound

2-Ethylhexanol (abbreviated 2-EH) is an organic compound with formula C8H18O. It is a branched, eight-carbon chiral alcohol. It is a colorless liquid that is poorly soluble in water but soluble in most organic solvents. It is produced on a large scale (>2,000,000,000 kg/y) for use in numerous applications such as solvents, flavors, and fragrances and especially as a precursor for production of other chemicals such as emollients and plasticizers. It is encountered in plants, fruits, and wines. The odor has been reported as "heavy, earthy, and slightly floral" for the R enantiomer and "a light, sweet floral fragrance" for the S enantiomer.

<span class="mw-page-title-main">Dimethyl carbonate</span> Chemical compound

Dimethyl carbonate (DMC) is an organic compound with the formula OC(OCH3)2. It is a colourless, flammable liquid. It is classified as a carbonate ester. This compound has found use as a methylating agent and as a co-solvent in lithium-ion batteries. Notably, dimethyl carbonate is a weak methylating agent, and is not considered as a carcinogen. Instead, dimethyl carbonate is often considered to be a green reagent, and it is exempt from the restrictions placed on most volatile organic compounds (VOCs) in the United States.

<span class="mw-page-title-main">Thermal oxidizer</span>

A thermal oxidizer is a process unit for air pollution control in many chemical plants that decomposes hazardous gases at a high temperature and releases them into the atmosphere.

<span class="mw-page-title-main">Air pollution</span> Presence of dangerous substances in the atmosphere

Air pollution is the contamination of air due to the presence of substances called pollutants in the atmosphere that are harmful to the health of humans and other living beings, or cause damage to the climate or to materials. It is also the contamination of the indoor or outdoor environment either by chemical, physical, or biological agents that alters the natural features of the atmosphere. There are many different types of air pollutants, such as gases, particulates, and biological molecules. Air pollution can cause diseases, allergies, and even death to humans; it can also cause harm to other living organisms such as animals and crops, and may damage the natural environment or built environment. Air pollution can be caused by both human activities and natural phenomena.

<span class="mw-page-title-main">Environmental effects of paint</span>

The environmental effects of paint can vary depending on the type of paint used and mitigation measures. Traditional painting materials and processes can have harmful effects on the environment, including those from the use of lead and other additives. Measures can be taken to reduce its environmental effects, including accurately estimating paint quantities so waste is minimized, and use of environmentally preferred paints, coating, painting accessories, and techniques.

<span class="mw-page-title-main">Environmental impact of the petroleum industry</span>

The environmental impact of the petroleum industry is extensive and expansive due to petroleum having many uses. Crude oil and natural gas are primary energy and raw material sources that enable numerous aspects of modern daily life and the world economy. Their supply has grown quickly over the last 150 years to meet the demands of the rapidly increasing human population, creativity, knowledge, and consumerism.

Organic molecular tracers in pollution control and environmental science are referred to as organic molecular markers or emission markers, and are compounds or compound classes. These tracers are of interest in the field of air quality because they can help identify particulate emission sources, as they are relatively unique to those sources. This approach is generally applied to particulate matter under 2.5μm in diameter because of the formation mechanisms and the health risks associated with this size regime. With tracer compounds, the principles of mass balance are used to 'trace' emissions from the source to the receptor site where a sample is taken. Use of organic tracers has become more common as measurement quality has improved, costs have decreased, and compounds that were historically good tracers, such as lead, have decreased in ambient concentrations due to various factors including government regulation.

Analytical thermal desorption, known within the analytical chemistry community simply as "thermal desorption" (TD), is a technique that concentrates volatile organic compounds (VOCs) in gas streams prior to injection into a gas chromatograph (GC). It can be used to lower the detection limits of GC methods, and can improve chromatographic performance by reducing peak widths.

Volatolomics is a branch of chemistry that studies volatile organic compounds (VOCs) emitted by a biological system, under specific experimental conditions.

<span class="mw-page-title-main">Floral scent</span>

Floral scent, or flower scent, is composed of all the volatile organic compounds (VOCs), or aroma compounds, emitted by floral tissue. Other names for floral scent include, aroma, fragrance, floral odour or perfume. Flower scent of most flowering plant species encompasses a diversity of VOCs, sometimes up to several hundred different compounds. The primary functions of floral scent are to deter herbivores and especially folivorous insects, and to attract pollinators. Floral scent is one of the most important communication channels mediating plant-pollinator interactions, along with visual cues.

<span class="mw-page-title-main">Flash-gas (petroleum)</span>

In an oil and gas production, flash-gas is a spontaneous vapor that is produced from the heating or depressurization of the extracted oil mixture during different phases of production. Flash evaporation, or flashing, is the process of volatile components suddenly vaporizing from their liquid state. This often happens during the transportation of petroleum products through pipelines and into vessels, such as when the stream from a common separation unit flows into an on-site atmospheric storage tank. Vessels that are used to intentionally “flash” a mixture of gas and saturated liquids are aptly named "flash drums." A type of vapor-liquid separator. A venting apparatus is used in these vessels to prevent damage due to increasing pressure, extreme cases of this are referred to as boiling liquid expanding vapor explosion (BLEVE).

References

  1. Carroll, Gregory T.; Kirschman, David L. (2022-12-20). "A Peripherally Located Air Recirculation Device Containing an Activated Carbon Filter Reduces VOC Levels in a Simulated Operating Room". ACS Omega. 7 (50): 46640–46645. doi:10.1021/acsomega.2c05570. ISSN   2470-1343. PMC   9774396 . PMID   36570243.
  2. 1 2 Koppmann, Ralf, ed. (2007). Volatile Organic Compounds in the Atmosphere. doi:10.1002/9780470988657. ISBN   9780470988657.
  3. Pichersky, Eran; Gershenzon, Jonathan (2002). "The formation and function of plant volatiles: Perfumes for pollinator attraction and defense". Current Opinion in Plant Biology. 5 (3): 237–243. doi:10.1016/S1369-5266(02)00251-0. PMID   11960742.
  4. Kessler, A.; Baldwin, I. T. (2001). "Defensive Function of Herbivore-Induced Plant Volatile Emissions in Nature". Science. 291 (5511): 2141–2144. Bibcode:2001Sci...291.2141K. doi:10.1126/science.291.5511.2141. PMID   11251117.
  5. Baldwin, I. T.; Halitschke, R.; Paschold, A.; von Dahl, C. C.; Preston, C. A. (2006). "Volatile Signaling in Plant-Plant Interactions: "Talking Trees" in the Genomics Era". Science. 311 (5762): 812–815. Bibcode:2006Sci...311..812B. doi:10.1126/science.1118446. PMID   16469918. S2CID   9260593.
  6. Health Canada Archived February 7, 2009, at the Wayback Machine
  7. Industrial Emissions Directive, article 3(45).
  8. The VOC solvent emission directive EUR-Lex, European Union Publications Office. Retrieved on 2010-09-28.
  9. The Paints Directive EUR-Lex, European Union Publications Office.
  10. eBeijing.gov.cn
  11. "国务院关于印发打赢蓝天保卫战三年行动计划的通知(国发〔2018〕22号)_政府信息公开专栏". www.gov.cn. Archived from the original on 2019-03-09.
  12. "THE AIR (PREVENTION AND CONTROL OF POLLUTION) ACT, 1981".
  13. "Air Pollution in IndiaClean Air India Movement". Clean Air India Movement.
  14. EPA. "Air Pollution Control Technology Fact Sheet: Thermal Incinerator." EPA-452/F-03-022.
  15. "CARB regulations on VOC in consumer products". Consumer Product Testing. Eurofins Scientific. 2016-08-19.
  16. "Definitions of VOC and ROG" (PDF). Sacramento, CA: California Air Resources Board. November 2004.
  17. For example, discharges from chemical and plastics manufacturing plants: "Organic Chemicals, Plastics and Synthetic Fibers Effluent Guidelines". EPA. 2016-02-01.
  18. Under the CERCLA ("Superfund") law and the Resource Conservation and Recovery Act.
  19. 1 2 "Volatile Organic Compounds' Impact on Indoor Air Quality". EPA. 2016-09-07.
  20. 1 2 Sindelarova, K.; Granier, C.; Bouarar, I.; Guenther, A.; Tilmes, S.; Stavrakou, T.; Müller, J.-F.; Kuhn, U.; Stefani, P.; Knorr, W. (2014). "Global data set of biogenic VOC emissions calculated by the MEGAN model over the last 30 years". Atmospheric Chemistry and Physics. 14 (17): 9317–9341. Bibcode:2014ACP....14.9317S. doi: 10.5194/acp-14-9317-2014 . hdl: 11858/00-001M-0000-0023-F4FB-B .
  21. J. Kesselmeier; M. Staudt (1999). "Biogenic Volatile Organic Compounds (VOC): An Overview on Emission, Physiology and Ecology". Journal of Atmospheric Chemistry. 33 (1): 23–88. Bibcode:1999JAtC...33...23K. doi:10.1023/A:1006127516791. S2CID   94021819.
  22. Terra, W. C.; Campos, V. P.; Martins, S. J. (2018). "Volatile organic molecules from Fusarium oxysporum strain 21 with nematicidal activity against Meloidogyne incognita". Crop Protection. 106: 125–131. doi:10.1016/j.cropro.2017.12.022.
  23. Kwak, Seon-Yeong; Wong, Min Hao; Lew, Tedrick Thomas Salim; Bisker, Gili; Lee, Michael A.; Kaplan, Amir; Dong, Juyao; Liu, Albert Tianxiang; Koman, Volodymyr B.; Sinclair, Rosalie; Hamann, Catherine; Strano, Michael S. (2017-06-12). "Nanosensor Technology Applied to Living Plant Systems". Annual Review of Analytical Chemistry . Annual Reviews. 10 (1): 113–140. doi:10.1146/annurev-anchem-061516-045310. ISSN   1936-1327. PMID   28605605.
  24. J. Lelieveld; T. M. Butler; J. N. Crowley; T. J. Dillon; H. Fischer; L. Ganzeveld; H. Harder; M. G. Lawrence; M. Martinez; D. Taraborrelli; J. Williams (2008). "Atmospheric oxidation capacity sustained by a tropical forest". Nature. 452 (7188): 737–740. Bibcode:2008Natur.452..737L. doi:10.1038/nature06870. PMID   18401407. S2CID   4341546.
  25. Josep Peñuelas; Michael Staudt (2010). "BVOCs and global change". Trends in Plant Science. 15 (3): 133–144. doi:10.1016/j.tplants.2009.12.005. PMID   20097116.
  26. Niinemets, Ülo; Loreto, Francesco; Reichstein, Markus (2004). "Physiological and physicochemical controls on foliar volatile organic compound emissions". Trends in Plant Science. 9 (4): 180–6. doi:10.1016/j.tplants.2004.02.006. PMID   15063868.
  27. Behr, Arno; Johnen, Leif (2009). "Myrcene as a Natural Base Chemical in Sustainable Chemistry: A Critical Review". ChemSusChem. 2 (12): 1072–95. doi:10.1002/cssc.200900186. PMID   20013989.
  28. Farag, Mohamed A.; Fokar, Mohamed; Abd, Haggag; Zhang, Huiming; Allen, Randy D.; Paré, Paul W. (2004). "(Z)-3-Hexenol induces defense genes and downstream metabolites in maize". Planta. 220 (6): 900–9. doi:10.1007/s00425-004-1404-5. PMID   15599762. S2CID   21739942.
  29. Goldstein, Allen H.; Galbally, Ian E. (2007). "Known and Unexplored Organic Constituents in the Earth's Atmosphere". Environmental Science & Technology. 41 (5): 1514–21. Bibcode:2007EnST...41.1514G. doi: 10.1021/es072476p . PMID   17396635.
  30. 1 2 Stefan Reimann; Alastair C. Lewis (2007). "Anthropogenic VOCs". In Koppmann, Ralf (ed.). Volatile Organic Compounds in the Atmosphere. doi:10.1002/9780470988657. ISBN   9780470988657.
  31. Stoye, D.; Funke, W.; Hoppe, L.; et al. (2006). "Paints and Coatings". Ullmann's Encyclopedia of Industrial Chemistry . Weinheim: Wiley-VCH. doi:10.1002/14356007.a18_359.pub2. ISBN   978-3527306732.
  32. Yeoman, Amber M.; Lewis, Alastair C. (2021-04-22). "Global emissions of VOCs from compressed aerosol products". Elementa: Science of the Anthropocene. 9 (1): 00177. doi: 10.1525/elementa.2020.20.00177 . ISSN   2325-1026.
  33. Jones, A.P. (1999). "Indoor air quality and health". Atmospheric Environment. 33 (28): 4535–64. Bibcode:1999AtmEn..33.4535J. doi:10.1016/S1352-2310(99)00272-1.
  34. 1 2 Wang, S.; Ang, H. M.; Tade, M. O. (2007). "Volatile organic compounds in indoor environment and photocatalytic oxidation: State of the art". Environment International. 33 (5): 694–705. doi:10.1016/j.envint.2007.02.011. PMID   17376530.
  35. 1 2 3 Holøs, S. B.; et al. (2019). "VOC emission rates in newly built and renovated buildings, and the influence of ventilation – a review and meta-analysis". Int. J. Of Ventilation. 18 (3): 153–166. doi:10.1080/14733315.2018.1435026. hdl: 10642/6247 . S2CID   56370102.
  36. 1 2 3 "KEY-VOCs". KEY-VOCs. Retrieved 23 April 2018.
  37. "ISO 16000-9:2006 Indoor air – Part 9: Determination of the emission of volatile organic compounds from building products and furnishing – Emission test chamber method". Iso.org. Retrieved 24 April 2018.
  38. Heeley-Hill, Aiden C.; Grange, Stuart K.; Ward, Martyn W.; Lewis, Alastair C.; Owen, Neil; Jordan, Caroline; Hodgson, Gemma; Adamson, Greg (2021). "Frequency of use of household products containing VOCs and indoor atmospheric concentrations in homes". Environmental Science: Processes & Impacts. 23 (5): 699–713. doi: 10.1039/D0EM00504E . ISSN   2050-7887. PMID   34037627.
  39. "emicode – Eurofins Scientific". Eurofins.com.
  40. "m1 – Eurofins Scientific". Eurofins.com.
  41. "blue-angel – Eurofins Scientific". Eurofins.com.
  42. "GuT-label". gut-prodis.eu.
  43. "Nordic Swan Ecolabel". nordic-ecolabel.org.
  44. "EU Ecolable homepage". ec.europa.eu.
  45. "www.indoor-air-comfort.com – Eurofins Scientific". Indoor-air-comfort.com.
  46. "cdph – Eurofins Scientific". Eurofins.com.
  47. Mendell, M. J. (2007). "Indoor residential chemical emissions as risk factors for respiratory and allergic effects in children: A review". Indoor Air. 17 (4): 259–77. doi: 10.1111/j.1600-0668.2007.00478.x . PMID   17661923.
  48. Wolkoff, P.; Wilkins, C. K.; Clausen, P. A.; Nielsen, G. D. (2006). "Organic compounds in office environments – sensory irritation, odor, measurements and the role of reactive chemistry". Indoor Air. 16 (1): 7–19. doi: 10.1111/j.1600-0668.2005.00393.x . PMID   16420493.
  49. "What is Smog?", Canadian Council of Ministers of the Environment, CCME.ca Archived September 28, 2011, at the Wayback Machine
  50. EPA,OAR, US (29 May 2015). "Basic Information about Ozone | US EPA". US EPA. Retrieved 2018-01-23.
  51. "Volatile Organic Compounds (VOCs) in Your Home – EH: Minnesota Department of Health". Health.state.mn.us. Retrieved 2018-01-23.
  52. US EPA, OAR (2014-08-18). "Volatile Organic Compounds' Impact on Indoor Air Quality". US EPA. Retrieved 2019-04-04.
  53. "Volatile Organic Compounds' Impact on Indoor Air Quality". EPA. 2017-04-19.
  54. Kim, Sung R.; Halden, Rolf U.; Buckley, Timothy J. (2007-03-01). "Volatile Organic Compounds in Human Milk: Methods and Measurements". Environmental Science & Technology. 41 (5): 1662–1667. Bibcode:2007EnST...41.1662K. doi:10.1021/es062362y. ISSN   0013-936X. PMID   17396657.
  55. Phillips, M; Greenberg, J; Awad, J (1994-11-01). "Metabolic and environmental origins of volatile organic compounds in breath". Journal of Clinical Pathology. 47 (11): 1052–1053. doi:10.1136/jcp.47.11.1052. ISSN   0021-9746. PMC   503075 . PMID   7829686.
  56. Cao, Fengmei; Qin, Pan; Lu, Shaoyong; He, Qi; Wu, Fengchang; Sun, Hongwen; Wang, Lei; Li, Linlin (December 2018). "Measurement of volatile organic compounds and associated risk assessments through ingestion and dermal routes in Dongjiang Lake, China". Ecotoxicology and Environmental Safety. 165: 645–653. doi:10.1016/j.ecoenv.2018.08.108. PMID   30243211. S2CID   52821729.
  57. 1 2 Ahn, Kangmo; Kim, Jihyun; Kim, Ji-Yun (February 2019). "Volatile Organic Compounds Dysregulate the Expression of Antimicrobial Peptides in Human Epidermal Keratinocytes". Journal of Allergy and Clinical Immunology. 143 (2): AB132. doi: 10.1016/j.jaci.2018.12.402 . S2CID   86509634.
  58. Bönisch, Ulrike; Böhme, Alexander; Kohajda, Tibor; Mögel, Iljana; Schütze, Nicole; von Bergen, Martin; Simon, Jan C.; Lehmann, Irina; Polte, Tobias (2012-07-03). Idzko, Marco (ed.). "Volatile Organic Compounds Enhance Allergic Airway Inflammation in an Experimental Mouse Model". PLOS ONE. 7 (7): e39817. Bibcode:2012PLoSO...739817B. doi: 10.1371/journal.pone.0039817 . ISSN   1932-6203. PMC   3389035 . PMID   22802943.
  59. Lee, Hana; Shin, Jung Jin; Bae, Hyun Cheol; Ryu, Woo-In; Son, Sang Wook (January 2017). "Toluene downregulates filaggrin expression via the extracellular signal-regulated kinase and signal transducer and activator of transcription-dependent pathways". Journal of Allergy and Clinical Immunology. 139 (1): 355–358.e5. doi: 10.1016/j.jaci.2016.06.036 . PMID   27498358.
  60. Huss-Marp, J.; Eberlein-Konig, B.; Breuer, K.; Mair, S.; Ansel, A.; Darsow, U.; Kramer, U.; Mayer, E.; Ring, J.; Behrendt, H. (March 2006). "Influence of short-term exposure to airborne Der p 1 and volatile organic compounds on skin barrier function and dermal blood flow in patients with atopic eczema and healthy individuals". Clinical & Experimental Allergy. 36 (3): 338–345. doi:10.1111/j.1365-2222.2006.02448.x. ISSN   0954-7894. PMID   16499645. S2CID   23522130.
  61. "Ausschuss zur gesundheitlichen Bewertung von Bauprodukten". Umweltbundesamt (in German). 2013-04-08. Retrieved 2019-05-24.
  62. EPA,OAR,ORIA,IED, US (18 August 2014). "Technical Overview of Volatile Organic Compounds | US EPA". US EPA. Retrieved 2018-04-23.{{cite web}}: CS1 maint: multiple names: authors list (link)
  63. "Before You Buy Paint". Consumer Information. 2012-10-09. Retrieved 2018-04-30.
  64. 1 2 Virji, M Abbas; Liang, Xiaoming; Su, Feng-Chiao; Lebouf, Ryan F; Stefaniak, Aleksandr B; Stanton, Marcia L; Henneberger, Paul K; Houseman, E Andres (2019-10-28). "Corrigendum to: Peaks, Means, and Determinants of Real-Time TVOC Exposures Associated with Cleaning and Disinfecting Tasks in Healthcare Settings". Annals of Work Exposures and Health. 64 (9): 1041. doi: 10.1093/annweh/wxz059 . ISSN   2398-7308. PMID   31665213.
  65. Charlier, Bruno; Coglianese, Albino; De Rosa, Federica; De Caro, Francesco; Piazza, Ornella; Motta, Oriana; Borrelli, Anna; Capunzo, Mario; Filippelli, Amelia; Izzo, Viviana (2021-03-24). "Chemical risk in hospital settings: Overview on monitoring strategies and international regulatory aspects". Journal of Public Health Research. 10 (1): jphr.2021.1993. doi:10.4081/jphr.2021.1993. ISSN   2279-9036. PMC   8018262 . PMID   33849259.
  66. 1 2 3 Su, Feng-Chiao; Friesen, Melissa C; Stefaniak, Aleksandr B; Henneberger, Paul K; LeBouf, Ryan F; Stanton, Marcia L; Liang, Xiaoming; Humann, Michael; Virji, M Abbas (2018-08-13). "Exposures to Volatile Organic Compounds among Healthcare Workers: Modeling the Effects of Cleaning Tasks and Product Use". Annals of Work Exposures and Health. 62 (7): 852–870. doi:10.1093/annweh/wxy055. ISSN   2398-7308. PMC   6248410 . PMID   29931140.
  67. 1 2 Bessonneau, Vincent; Mosqueron, Luc; Berrubé, Adèle; Mukensturm, Gaël; Buffet-Bataillon, Sylvie; Gangneux, Jean-Pierre; Thomas, Olivier (2013-02-05). Levin, Jan-Olof (ed.). "VOC Contamination in Hospital, from Stationary Sampling of a Large Panel of Compounds, in View of Healthcare Workers and Patients Exposure Assessment". PLOS ONE. 8 (2): e55535. Bibcode:2013PLoSO...855535B. doi: 10.1371/journal.pone.0055535 . ISSN   1932-6203. PMC   3564763 . PMID   23393590.
  68. 1 2 LeBouf, Ryan F; Virji, M Abbas; Saito, Rena; Henneberger, Paul K; Simcox, Nancy; Stefaniak, Aleksandr B (September 2014). "Exposure to volatile organic compounds in healthcare settings". Occupational and Environmental Medicine. 71 (9): 642–650. doi:10.1136/oemed-2014-102080. ISSN   1351-0711. PMC   4591534 . PMID   25011549.
  69. 1 2 Reddy, Manasa; Heidarinejad, Mohammad; Stephens, Brent; Rubinstein, Israel (April 2021). "Adequate indoor air quality in nursing homes: An unmet medical need". Science of the Total Environment. 765: 144273. Bibcode:2021ScTEn.765n4273R. doi:10.1016/j.scitotenv.2020.144273. PMID   33401060. S2CID   230782257.
  70. Belo, Joana; Carreiro-Martins, Pedro; Papoila, Ana L.; Palmeiro, Teresa; Caires, Iolanda; Alves, Marta; Nogueira, Susana; Aguiar, Fátima; Mendes, Ana; Cano, Manuela; Botelho, Maria A. (2019-10-15). "The impact of indoor air quality on respiratory health of older people living in nursing homes: spirometric and exhaled breath condensate assessments". Journal of Environmental Science and Health, Part A. 54 (12): 1153–1158. doi:10.1080/10934529.2019.1637206. ISSN   1093-4529. PMID   31274053. S2CID   195807320.
  71. Bentayeb, Malek; Billionnet, Cécile; Baiz, Nour; Derbez, Mickaël; Kirchner, Séverine; Annesi-Maesano, Isabella (October 2013). "Higher prevalence of breathlessness in elderly exposed to indoor aldehydes and VOCs in a representative sample of French dwellings". Respiratory Medicine. 107 (10): 1598–1607. doi: 10.1016/j.rmed.2013.07.015 . PMID   23920330.
  72. Lattuati-Derieux, Agnès; Bonnassies-Termes, Sylvette; Lavédrine, Bertrand (2004). "Identification of volatile organic compounds emitted by a naturally aged book using solid-phase microextraction/gas chromatography/mass spectrometry". Journal of Chromatography A. 1026 (1–2): 9–18. doi:10.1016/j.chroma.2003.11.069. PMID   14870711.
  73. 1 2 Ahmed, Waqar M.; Lawal, Oluwasola; Nijsen, Tamara M.; Goodacre, Royston; Fowler, Stephen J. (2017). "Exhaled Volatile Organic Compounds of Infection: A Systematic Review". ACS Infectious Diseases. 3 (10): 695–710. doi:10.1021/acsinfecdis.7b00088. PMID   28870074.
  74. Who Says Alcohol and Benzene Don't Mix? Archived April 15, 2008, at the Wayback Machine
  75. Fang, Shuting; Liu, Shuqin; Song, Juyi; Huang, Qihong; Xiang, Zhangmin (2021-04-01). "Recognition of pathogens in food matrixes based on the untargeted in vivo microbial metabolite profiling via a novel SPME/GC × GC-QTOFMS approach". Food Research International. 142: 110213. doi:10.1016/j.foodres.2021.110213. ISSN   0963-9969. PMID   33773687. S2CID   232407164.
  76. Biasioli, Franco; Yeretzian, Chahan; Märk, Tilmann D.; Dewulf, Jeroen; Van Langenhove, Herman (2011). "Direct-injection mass spectrometry adds the time dimension to (B)VOC analysis". Trends in Analytical Chemistry. 30 (7): 1003–1017. doi:10.1016/j.trac.2011.04.005.
  77. Ellis, Andrew M.; Mayhew, Christopher A. (2014). Proton Transfer Reaction Mass Spectrometry – Principles and Applications. Chichester, West Sussex, UK: John Wiley & Sons Ltd. ISBN   978-1-405-17668-2.
  78. Sulzer, Philipp; Hartungen, Eugen; Hanel, Gernot; Feil, Stefan; Winkler, Klaus; Mutschlechner, Paul; Haidacher, Stefan; Schottkowsky, Ralf; Gunsch, Daniel; Seehauser, Hans; Striednig, Marcus; Jürschik, Simone; Breiev, Kostiantyn; Lanza, Matteo; Herbig, Jens; Märk, Lukas; Märk, Tilmann D.; Jordan, Alfons (2014). "A Proton Transfer Reaction-Quadrupole interface Time-Of-Flight Mass Spectrometer (PTR-QiTOF): High speed due to extreme sensitivity". International Journal of Mass Spectrometry. 368: 1–5. Bibcode:2014IJMSp.368....1S. doi:10.1016/j.ijms.2014.05.004.
  79. 1 2 Buszewski, B. A.; et al. (2007). "Human exhaled air analytics: Biomarkers of diseases". Biomedical Chromatography. 21 (6): 553–566. doi: 10.1002/bmc.835 . PMID   17431933.
  80. Miekisch, W.; Schubert, J. K.; Noeldge-Schomburg, G. F. E. (2004). "Diagnostic potential of breath analysis—focus on volatile organic compounds". Clinica Chimica Acta. 347 (1–2): 25–39. doi:10.1016/j.cccn.2004.04.023. PMID   15313139.
  81. Mazzone, P. J. (2008). "Analysis of Volatile Organic Compounds in the Exhaled Breath for the Diagnosis of Lung Cancer". Journal of Thoracic Oncology. 3 (7): 774–780. doi: 10.1097/JTO.0b013e31817c7439 . PMID   18594325.
  82. Hoerger, C. C.; Claude, A., Plass-Duelmer, C., Reimann, S., Eckart, E., Steinbrecher, R., Aalto, J., Arduini, J., Bonnaire, N., Cape, J. N., Colomb, A., Connolly, R., Diskova, J., Dumitrean, P., Ehlers, C., Gros, V., Hakola, H., Hill, M., Hopkins, J. R., Jäger, J., Junek, R., Kajos, M. K., Klemp, D., Leuchner, M., Lewis, A. C., Locoge, N., Maione, M., Martin, D., Michl, K., Nemitz, E., O'Doherty, S., Pérez Ballesta, P., Ruuskanen, T. M., Sauvage, S., Schmidbauer, N., Spain, T. G., Straube, E., Vana, M., Vollmer, M. K., Wegener, R., Wenger, A. (2015). "ACTRIS non-methane hydrocarbon intercomparison experiment in Europe to support WMO GAW and EMEP observation networks". Atmospheric Measurement Techniques. 8 (7): 2715–2736. Bibcode:2015AMT.....8.2715H. doi: 10.5194/amt-8-2715-2015 . hdl: 1983/f9d95320-dcc6-48d1-a58a-bf310a536b9c .{{cite journal}}: CS1 maint: multiple names: authors list (link)