What elements make up carbon dioxide

Carbon dioxide is a tasteless, odorless gas that makes up less than 1 percent of the Earth’s atmosphere. It’s a natural product of respiration, and plants use it to make food. Too much carbon dioxide is dangerous. It contributes to global warming, and if you can’t breathe out the carbon dioxide your body makes, you can become seriously ill or die.

Chemical Makeup

A molecule of the compound carbon dioxide contains one atom of the element carbon and two atoms of the element oxygen. Each oxygen atom shares a double bond with the carbon atom. Carbon is the sixth element in the periodic table and occurs in pure form as coal and diamonds. Oxygen, element eight, is a gas that makes up about 21 percent of the atmosphere. Living things take in oxygen and release carbon dioxide. The oxygen helps turn food into energy, creating carbon dioxide as a byproduct. Green plants use carbon dioxide in photosynthesis to make and store carbohydrates, the plants’ food source.

Carbon Dioxide Dangers

Hypercapnia is a disorder in which too much carbon dioxide builds up in the bloodstream. It can be fatal and is one of the causes of sudden infant death syndrome, or SIDS. The causes of hypercapnia include lung disease, rebreathing exhaled air and exposure to high concentrations of carbon dioxide. As a greenhouse gas, carbon dioxide contributes to global warming, which, if left unchecked, might create many problems for planet Earth.

References

Resources

Writer Bio

Based in Greenville SC, Eric Bank has been writing business-related articles since 1985. He holds an M.B.A. from New York University and an M.S. in finance from DePaul University. You can see samples of his work at ericbank.com.

Carbon dioxide is a very prevalent molecule. It is a product of respiration in humans and other animals, and green plants use carbon dioxide and water to form carbohydrates in photosynthesis. Carbon dioxide emissions, produced when any carbon-containing substance is burned, are a significant contributor to global climate change. It is also used in refrigeration and for beverage carbonation.

Anatomy of a Greenhouse Gas

The carbon dioxide molecule contains one carbon and two oxygen atoms. The molecule is linear, with the carbon atom in the center, forming a double bond with an oxygen on each side. Carbon dioxide is an odorless, colorless, nonflammable gas at room temperature. It exists as a solid at negative 78 degrees Celsius (negative 108.4 degrees Fahrenheit). In this form it is commonly known as dry ice. Carbon dioxide is water-soluble when pressure is sufficiently high. Once pressure drops, carbon dioxide will try to escape, forming bubbles that are recognizable as carbonation.

Carbon dioxide

What elements make up carbon dioxide

What elements make up carbon dioxide

What elements make up carbon dioxide

Names
Other names

  • Carbonic acid gas
  • Carbonic anhydride
  • Carbonic dioxide
  • Carbon(IV) oxide
  • R-744 (refrigerant)
  • R744 (refrigerant alternative spelling)
  • Dry ice (solid phase)

Identifiers

CAS Number

  • 124-38-9 
    What elements make up carbon dioxide

3D model (JSmol)

  • Interactive image
  • Interactive image

3DMet

  • B01131

Beilstein Reference

1900390
ChEBI

  • CHEBI:16526 
    What elements make up carbon dioxide

ChEMBL

  • ChEMBL1231871 
    What elements make up carbon dioxide

ChemSpider

  • 274 
    What elements make up carbon dioxide

ECHA InfoCard 100.004.271
What elements make up carbon dioxide
EC Number

  • 204-696-9

E number E290 (preservatives)

Gmelin Reference

989
KEGG

  • D00004 
    What elements make up carbon dioxide

MeSH Carbon+dioxide

PubChem CID

  • 280

RTECS number

  • FF6400000

UNII

  • 142M471B3J 
    What elements make up carbon dioxide

UN number 1013 (gas), 1845 (solid)

CompTox Dashboard (EPA)

  • DTXSID4027028
    What elements make up carbon dioxide

InChI

  • InChI=1S/CO2/c2-1-3 

    What elements make up carbon dioxide

    Key: CURLTUGMZLYLDI-UHFFFAOYSA-N 

    What elements make up carbon dioxide

  • InChI=1/CO2/c2-1-3

    Key: CURLTUGMZLYLDI-UHFFFAOYAO

SMILES

  • O=C=O

  • C(=O)=O

Properties

Chemical formula

CO2
Molar mass 44.009 g·mol−1
Appearance Colorless gas
Odor

  • Low concentrations: none
  • High concentrations: sharp; acidic[1]

Density

  • 1562 kg/m3 (solid at 1 atm (100 kPa) and −78.5 °C (−109.3 °F))
  • 1101 kg/m3 (liquid at saturation −37 °C (−35 °F))
  • 1.977 kg/m3 (gas at 1 atm (100 kPa) and 0 °C (32 °F))

Critical point (T, P) 304.128(15) K[2] (30.978(15) °C), 7.3773(30) MPa[2] (72.808(30) atm)

Sublimation
conditions

194.6855(30) K (−78.4645(30) °C) at 1 atm (0.101325 MPa)

Solubility in water

1.45 g/L at 25 °C (77 °F), 100 kPa (0.99 atm)
Vapor pressure 5.7292(30) MPa, 56.54(30) atm (20 °C (293.15 K))
Acidity (pKa) 6.35, 10.33

Magnetic susceptibility (χ)

−20.5·10−6 cm3/mol
Thermal conductivity 0.01662 W·m−1·K−1 (300 K (27 °C; 80 °F))[3]

Refractive index (nD)

1.00045
Viscosity

  • 14.90 μPa·s at 25 °C (298 K)[4]
  • 70 μPa·s at −78.5 °C (194.7 K)

Dipole moment

0 D
Structure

Crystal structure

Trigonal

Molecular shape

Linear
Thermochemistry

Heat capacity (C)

37.135 J/K·mol

Std molar
entropy (S⦵298)

214 J·mol−1·K−1

Std enthalpy of
formation (ΔfH⦵298)

−393.5 kJ·mol−1
Pharmacology

ATC code

V03AN02 (WHO)
Hazards
NFPA 704 (fire diamond)

[7][8]

What elements make up carbon dioxide

2

0

0

SA

Lethal dose or concentration (LD, LC):

LCLo (lowest published)

90,000 ppm (human, 5 min)[6]
NIOSH (US health exposure limits):

PEL (Permissible)

TWA 5000 ppm (9000 mg/m3)[5]

REL (Recommended)

TWA 5000 ppm (9000 mg/m3), ST 30,000 ppm (54,000 mg/m3)[5]

IDLH (Immediate danger)

40,000 ppm[5]
Safety data sheet (SDS) Sigma-Aldrich
Related compounds

Other anions

  • Carbon disulfide
  • Carbon diselenide
  • Carbon ditelluride

Other cations

  • Silicon dioxide
  • Germanium dioxide
  • Tin dioxide
  • Lead dioxide

Related carbon oxides

  • Carbon monoxide
  • Carbon suboxide
  • Dicarbon monoxide
  • Carbon trioxide

Related compounds

  • Carbonic acid
  • Carbonyl sulfide

Supplementary data page
Carbon dioxide (data page)

Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).

What elements make up carbon dioxide
 verify (what is 
What elements make up carbon dioxide
What elements make up carbon dioxide
 ?)

Infobox references

Carbon dioxide (chemical formula CO2) is a chemical compound made up of molecules that each have one carbon atom covalently double bonded to two oxygen atoms. It is found in the gas state at room temperature.

In the air, carbon dioxide is transparent to visible light but absorbs infrared radiation, acting as a greenhouse gas. It is a trace gas in Earth's atmosphere at 417 ppm (about 0.04%) by volume, having risen from pre-industrial levels of 280 ppm.[9][10] Burning fossil fuels is the primary cause of these increased CO2 concentrations and also the primary cause of global warming and climate change.[11] Carbon dioxide is soluble in water and is found in groundwater, lakes, ice caps, and seawater. When carbon dioxide dissolves in water it forms carbonic acid (H2CO3), which causes ocean acidification as atmospheric CO2 levels increase.[12]

As the source of available carbon in the carbon cycle, atmospheric carbon dioxide is the primary carbon source for life on Earth. Its concentration in Earth's pre-industrial atmosphere since late in the Precambrian has been regulated by organisms and geological phenomena. Plants, algae and cyanobacteria use energy from sunlight to synthesize carbohydrates from carbon dioxide and water in a process called photosynthesis, which produces oxygen as a waste product.[13] In turn, oxygen is consumed and CO2 is released as waste by all aerobic organisms when they metabolize organic compounds to produce energy by respiration.[14] CO2 is released from organic materials when they decay or combust, such as in forest fires. Since plants require CO2 for photosynthesis, and humans and animals depend on plants for food, CO2 is necessary for the survival of life on earth.

Carbon dioxide is 53% more dense than dry air, but is long lived and thoroughly mixes in the atmosphere. About half of excess CO2 emissions to the atmosphere are absorbed by land and ocean carbon sinks.[15] These sinks can become saturated and are volatile, as decay and wildfires result in the CO2 being released back into the atmosphere.[16] CO2 is eventually sequestered (stored for the long term) in rocks and organic deposits like coal, petroleum and natural gas. Sequestered CO2 is released into the atmosphere through burning fossil fuels or naturally by volcanoes, hot springs, geysers, and when carbonate rocks dissolve in water or react with acids.

CO2 is a versatile industrial material, used, for example, as an inert gas in welding and fire extinguishers, as a pressurizing gas in air guns and oil recovery, and as a supercritical fluid solvent in decaffeination of coffee and supercritical drying.[17] It is also a feedstock for the synthesis of fuels and chemicals.[18][19][20][21] It is an unwanted byproduct in many large scale oxidation processes, for example, in the production of acrylic acid (over 5 million tons/year).[22][23][24] The frozen solid form of CO2, known as dry ice, is used as a refrigerant and as an abrasive in dry-ice blasting. It is a byproduct of fermentation of sugars in bread, beer and wine making, and is added to carbonated beverages like seltzer and beer for effervescence. It has a sharp and acidic odor and generates the taste of soda water in the mouth,[25] but at normally encountered concentrations it is odorless.[1]

History

What elements make up carbon dioxide

Carbon dioxide was the first gas to be described as a discrete substance. In about 1640,[26] the Flemish chemist Jan Baptist van Helmont observed that when he burned charcoal in a closed vessel, the mass of the resulting ash was much less than that of the original charcoal. His interpretation was that the rest of the charcoal had been transmuted into an invisible substance he termed a "gas" or "wild spirit" (spiritus sylvestris).[27]

The properties of carbon dioxide were further studied in the 1750s by the Scottish physician Joseph Black. He found that limestone (calcium carbonate) could be heated or treated with acids to yield a gas he called "fixed air." He observed that the fixed air was denser than air and supported neither flame nor animal life. Black also found that when bubbled through limewater (a saturated aqueous solution of calcium hydroxide), it would precipitate calcium carbonate. He used this phenomenon to illustrate that carbon dioxide is produced by animal respiration and microbial fermentation. In 1772, English chemist Joseph Priestley published a paper entitled Impregnating Water with Fixed Air in which he described a process of dripping sulfuric acid (or oil of vitriol as Priestley knew it) on chalk in order to produce carbon dioxide, and forcing the gas to dissolve by agitating a bowl of water in contact with the gas.[28]

Carbon dioxide was first liquefied (at elevated pressures) in 1823 by Humphry Davy and Michael Faraday.[29] The earliest description of solid carbon dioxide (dry ice) was given by the French inventor Adrien-Jean-Pierre Thilorier, who in 1835 opened a pressurized container of liquid carbon dioxide, only to find that the cooling produced by the rapid evaporation of the liquid yielded a "snow" of solid CO2.[30][31]

Chemical and physical properties

Structure, bonding and molecular vibrations

The symmetry of a carbon dioxide molecule is linear and centrosymmetric at its equilibrium geometry. The length of the carbon-oxygen bond in carbon dioxide is 116.3 pm, noticeably shorter than the roughly 140-pm length of a typical single C–O bond, and shorter than most other C–O multiply-bonded functional groups such as carbonyls.[32] Since it is centrosymmetric, the molecule has no electric dipole moment.

What elements make up carbon dioxide

Stretching and bending oscillations of the CO2 carbon dioxide molecule. Upper left: symmetric stretching. Upper right: antisymmetric stretching. Lower line: degenerate pair of bending modes.

As a linear triatomic molecule, CO2 has four vibrational modes as shown in the diagram. In the symmetric and the antisymmetric stretching modes, the atoms move along the axis of the molecule. There are two bending modes, which are degenerate, meaning that they have the same frequency and same energy, because of the symmetry of the molecule. When a molecule touches a surface or touches another molecule, the two bending modes can differ in frequency because the interaction is different for the two modes. Some of the vibrational modes are observed in the infrared (IR) spectrum: the antisymmetric stretching mode at wavenumber 2349 cm−1 (wavelength 4.25 μm) and the degenerate pair of bending modes at 667 cm−1 (wavelength 15 μm). The symmetric stretching mode does not create an electric dipole so is not observed in IR spectroscopy, but it is detected in by Raman spectroscopy at 1388 cm−1 (wavelength 7.2 μm).[33]

In the gas phase, carbon dioxide molecules undergo significant vibrational motions and do not keep a fixed structure. However, in a Coulomb explosion imaging experiment, an instantaneous image of the molecular structure can be deduced. Such an experiment[34] has been performed for carbon dioxide. The result of this experiment, and the conclusion of theoretical calculations[35] based on an ab initio potential energy surface of the molecule, is that none of the molecules in the gas phase are ever exactly linear.

In aqueous solution

Carbon dioxide is soluble in water, in which it reversibly forms H2CO3 (carbonic acid), which is a weak acid since its ionization in water is incomplete.

The hydration equilibrium constant of carbonic acid is, at 25 °C:

Hence, the majority of the carbon dioxide is not converted into carbonic acid, but remains as CO2 molecules, not affecting the pH.

The relative concentrations of CO2, H2CO3, and the deprotonated forms HCO3 (bicarbonate) and CO2−3(carbonate) depend on the pH. As shown in a Bjerrum plot, in neutral or slightly alkaline water (pH > 6.5), the bicarbonate form predominates (>50%) becoming the most prevalent (>95%) at the pH of seawater. In very alkaline water (pH > 10.4), the predominant (>50%) form is carbonate. The oceans, being mildly alkaline with typical pH = 8.2–8.5, contain about 120 mg of bicarbonate per liter.

Being diprotic, carbonic acid has two acid dissociation constants, the first one for the dissociation into the bicarbonate (also called hydrogen carbonate) ion (HCO3):

Ka1 = 2.5×10−4 mol/L; pKa1 = 3.6 at 25 °C.[32]

This is the true first acid dissociation constant, defined as

where the denominator includes only covalently bound H2CO3 and does not include hydrated CO2(aq). The much smaller and often-quoted value near 4.16×10−7 is an apparent value calculated on the (incorrect) assumption that all dissolved CO2 is present as carbonic acid, so that

Since most of the dissolved CO2remains as CO2 molecules, Ka1(apparent) has a much larger denominator and a much smaller value than the true Ka1.[36]

The bicarbonate ion is an amphoteric species that can act as an acid or as a base, depending on pH of the solution. At high pH, it dissociates significantly into the carbonate ion (CO2−3):

Ka2 = 4.69×10−11 mol/L; pKa2 = 10.329

In organisms carbonic acid production is catalysed by the enzyme, carbonic anhydrase.

Chemical reactions of CO2

CO2 is a potent electrophile having an electrophilic reactivity that is comparable to benzaldehyde or strong α,β-unsaturated carbonyl compounds. However, unlike electrophiles of similar reactivity, the reactions of nucleophiles with CO2 are thermodynamically less favored and are often found to be highly reversible.[37] Only very strong nucleophiles, like the carbanions provided by Grignard reagents and organolithium compounds react with CO2 to give carboxylates:

where M = Li or Mg Br and R = alkyl or aryl.

In metal carbon dioxide complexes, CO2 serves as a ligand, which can facilitate the conversion of CO2 to other chemicals.[38]

The reduction of CO2 to CO is ordinarily a difficult and slow reaction:

Photoautotrophs (i.e. plants and cyanobacteria) use the energy contained in sunlight to photosynthesize simple sugars from CO2 absorbed from the air and water:

The redox potential for this reaction near pH 7 is about −0.53 V versus the standard hydrogen electrode. The nickel-containing enzyme carbon monoxide dehydrogenase catalyses this process.[39]

Physical properties

What elements make up carbon dioxide

Pellets of "dry ice", a common form of solid carbon dioxide

Carbon dioxide is colorless. At low concentrations the gas is odorless; however, at sufficiently high concentrations, it has a sharp, acidic odor.[1] At standard temperature and pressure, the density of carbon dioxide is around 1.98 kg/m3, about 1.53 times that of air.[40]

Carbon dioxide has no liquid state at pressures below 0.51795(10) MPa[2] (5.11177(99) atm). At a pressure of 1 atm (0.101325 MPa), the gas deposits directly to a solid at temperatures below 194.6855(30) K[2] (−78.4645(30) °C) and the solid sublimes directly to a gas above this temperature. In its solid state, carbon dioxide is commonly called dry ice.

What elements make up carbon dioxide

Pressure–temperature phase diagram of carbon dioxide. Note that it is a log-lin chart.

Liquid carbon dioxide forms only at pressures above 0.51795(10) MPa[2] (5.11177(99) atm); the triple point of carbon dioxide is 216.592(3) K[2] (−56.558(3) °C) at 0.51795(10) MPa[2] (5.11177(99) atm) (see phase diagram). The critical point is 304.128(15) K[2] (30.978(15) °C) at 7.3773(30) MPa[2] (72.808(30) atm). Another form of solid carbon dioxide observed at high pressure is an amorphous glass-like solid.[41] This form of glass, called carbonia, is produced by supercooling heated CO2 at extreme pressures (40–48 GPa, or about 400,000 atmospheres) in a diamond anvil. This discovery confirmed the theory that carbon dioxide could exist in a glass state similar to other members of its elemental family, like silicon dioxide (silica glass) and germanium dioxide. Unlike silica and germania glasses, however, carbonia glass is not stable at normal pressures and reverts to gas when pressure is released.

At temperatures and pressures above the critical point, carbon dioxide behaves as a supercritical fluid known as supercritical carbon dioxide.

Isolation and production

Carbon dioxide can be obtained by distillation from air, but the method is inefficient. Industrially, carbon dioxide is predominantly an unrecovered waste product, produced by several methods which may be practiced at various scales.[42]

The combustion of all carbon-based fuels, such as methane (natural gas), petroleum distillates (gasoline, diesel, kerosene, propane), coal, wood and generic organic matter produces carbon dioxide and, except in the case of pure carbon, water. As an example, the chemical reaction between methane and oxygen:

Iron is reduced from its oxides with coke in a blast furnace, producing pig iron and carbon dioxide:[43]

Carbon dioxide is a byproduct of the industrial production of hydrogen by steam reforming and the water gas shift reaction in ammonia production. These processes begin with the reaction of water and natural gas (mainly methane).[44] This is a major source of food-grade carbon dioxide for use in carbonation of beer and soft drinks, and is also used for stunning animals such as poultry. In the summer of 2018 a shortage of carbon dioxide for these purposes arose in Europe due to the temporary shut-down of several ammonia plants for maintenance.[45]

Carbonates

It is produced by thermal decomposition of limestone, CaCO
3
by heating (calcining) at about 850 °C (1,560 °F), in the manufacture of quicklime (calcium oxide, CaO), a compound that has many industrial uses:

Acids liberate CO2 from most metal carbonates. Consequently, it may be obtained directly from natural carbon dioxide springs, where it is produced by the action of acidified water on limestone or dolomite. The reaction between hydrochloric acid and calcium carbonate (limestone or chalk) is shown below:

The carbonic acid (H
2
CO
3
) then decomposes to water and CO2:

Such reactions are accompanied by foaming or bubbling, or both, as the gas is released. They have widespread uses in industry because they can be used to neutralize waste acid streams.

Fermentation

Carbon dioxide is a by-product of the fermentation of sugar in the brewing of beer, whisky and other alcoholic beverages and in the production of bioethanol. Yeast metabolizes sugar to produce CO2 and ethanol, also known as alcohol, as follows:

All aerobic organisms produce CO2 when they oxidize carbohydrates, fatty acids, and proteins. The large number of reactions involved are exceedingly complex and not described easily. Refer to (cellular respiration, anaerobic respiration and photosynthesis). The equation for the respiration of glucose and other monosaccharides is:

Anaerobic organisms decompose organic material producing methane and carbon dioxide together with traces of other compounds.[46] Regardless of the type of organic material, the production of gases follows well defined kinetic pattern. Carbon dioxide comprises about 40–45% of the gas that emanates from decomposition in landfills (termed "landfill gas"). Most of the remaining 50–55% is methane.[47]

Applications

Carbon dioxide is used by the food industry, the oil industry, and the chemical industry.[42] The compound has varied commercial uses but one of its greatest uses as a chemical is in the production of carbonated beverages; it provides the sparkle in carbonated beverages such as soda water, beer and sparkling wine.

Precursor to chemicals

What elements make up carbon dioxide

This section needs expansion. You can help by adding to it. (July 2014)

In the chemical industry, carbon dioxide is mainly consumed as an ingredient in the production of urea, with a smaller fraction being used to produce methanol and a range of other products.[48] Some carboxylic acid derivatives such as sodium salicylate are prepared using CO2 by the Kolbe-Schmitt reaction.[49]

In addition to conventional processes using CO2 for chemical production, electrochemical methods are also being explored at a research level. In particular, the use of renewable energy for production of fuels from CO2 (such as methanol) is attractive as this could result in fuels that could be easily transported and used within conventional combustion technologies but have no net CO2 emissions.[50]

Agriculture

Plants require carbon dioxide to conduct photosynthesis. The atmospheres of greenhouses may (if of large size, must) be enriched with additional CO2 to sustain and increase the rate of plant growth.[51][52] At very high concentrations (100 times atmospheric concentration, or greater), carbon dioxide can be toxic to animal life, so raising the concentration to 10,000 ppm (1%) or higher for several hours will eliminate pests such as whiteflies and spider mites in a greenhouse.[53]

Foods

What elements make up carbon dioxide

Carbon dioxide bubbles in a soft drink

Carbon dioxide is a food additive used as a propellant and acidity regulator in the food industry. It is approved for usage in the EU[54] (listed as E number E290), US[55] and Australia and New Zealand[56] (listed by its INS number 290).

A candy called Pop Rocks is pressurized with carbon dioxide gas[57] at about 4,000 kPa (40 bar; 580 psi). When placed in the mouth, it dissolves (just like other hard candy) and releases the gas bubbles with an audible pop.

Leavening agents cause dough to rise by producing carbon dioxide.[58] Baker's yeast produces carbon dioxide by fermentation of sugars within the dough, while chemical leaveners such as baking powder and baking soda release carbon dioxide when heated or if exposed to acids.

Beverages

Carbon dioxide is used to produce carbonated soft drinks and soda water. Traditionally, the carbonation of beer and sparkling wine came about through natural fermentation, but many manufacturers carbonate these drinks with carbon dioxide recovered from the fermentation process. In the case of bottled and kegged beer, the most common method used is carbonation with recycled carbon dioxide. With the exception of British real ale, draught beer is usually transferred from kegs in a cold room or cellar to dispensing taps on the bar using pressurized carbon dioxide, sometimes mixed with nitrogen.

The taste of soda water (and related taste sensations in other carbonated beverages) is an effect of the dissolved carbon dioxide rather than the bursting bubbles of the gas. Carbonic anhydrase 4 converts to carbonic acid leading to a sour taste, and also the dissolved carbon dioxide induces a somatosensory response.[59]

Winemaking

What elements make up carbon dioxide

Dry ice used to preserve grapes after harvest

Carbon dioxide in the form of dry ice is often used during the cold soak phase in winemaking to cool clusters of grapes quickly after picking to help prevent spontaneous fermentation by wild yeast. The main advantage of using dry ice over water ice is that it cools the grapes without adding any additional water that might decrease the sugar concentration in the grape must, and thus the alcohol concentration in the finished wine. Carbon dioxide is also used to create a hypoxic environment for carbonic maceration, the process used to produce Beaujolais wine.

Carbon dioxide is sometimes used to top up wine bottles or other storage vessels such as barrels to prevent oxidation, though it has the problem that it can dissolve into the wine, making a previously still wine slightly fizzy. For this reason, other gases such as nitrogen or argon are preferred for this process by professional wine makers.

Stunning animals

Carbon dioxide is often used to "stun" animals before slaughter.[60] "Stunning" may be a misnomer, as the animals are not knocked out immediately and may suffer distress.[61][62]

Inert gas

Carbon dioxide is one of the most commonly used compressed gases for pneumatic (pressurized gas) systems in portable pressure tools. Carbon dioxide is also used as an atmosphere for welding, although in the welding arc, it reacts to oxidize most metals. Use in the automotive industry is common despite significant evidence that welds made in carbon dioxide are more brittle than those made in more inert atmospheres.[citation needed] When used for MIG welding, CO2 use is sometimes referred to as MAG welding, for Metal Active Gas, as CO2 can react at these high temperatures. It tends to produce a hotter puddle than truly inert atmospheres, improving the flow characteristics. Although, this may be due to atmospheric reactions occurring at the puddle site. This is usually the opposite of the desired effect when welding, as it tends to embrittle the site, but may not be a problem for general mild steel welding, where ultimate ductility is not a major concern.

Carbon dioxide is used in many consumer products that require pressurized gas because it is inexpensive and nonflammable, and because it undergoes a phase transition from gas to liquid at room temperature at an attainable pressure of approximately 60 bar (870 psi; 59 atm), allowing far more carbon dioxide to fit in a given container than otherwise would. Life jackets often contain canisters of pressured carbon dioxide for quick inflation. Aluminium capsules of CO2 are also sold as supplies of compressed gas for air guns, paintball markers/guns, inflating bicycle tires, and for making carbonated water. High concentrations of carbon dioxide can also be used to kill pests. Liquid carbon dioxide is used in supercritical drying of some food products and technological materials, in the preparation of specimens for scanning electron microscopy[63] and in the decaffeination of coffee beans.

Fire extinguisher

What elements make up carbon dioxide

Use of a CO2 fire extinguisher

Carbon dioxide can be used to extinguish flames by flooding the environment around the flame with the gas. It does not itself react to extinguish the flame, but starves the flame of oxygen by displacing it. Some fire extinguishers, especially those designed for electrical fires, contain liquid carbon dioxide under pressure. Carbon dioxide extinguishers work well on small flammable liquid and electrical fires, but not on ordinary combustible fires, because they do not cool the burning substances significantly, and when the carbon dioxide disperses, they can catch fire upon exposure to atmospheric oxygen. They are mainly used in server rooms.[64]

Carbon dioxide has also been widely used as an extinguishing agent in fixed fire-protection systems for local application of specific hazards and total flooding of a protected space.[65] International Maritime Organization standards recognize carbon-dioxide systems for fire protection of ship holds and engine rooms. Carbon-dioxide-based fire-protection systems have been linked to several deaths, because it can cause suffocation in sufficiently high concentrations. A review of CO2 systems identified 51 incidents between 1975 and the date of the report (2000), causing 72 deaths and 145 injuries.[66]

Supercritical CO2 as solvent

Liquid carbon dioxide is a good solvent for many lipophilic organic compounds and is used to remove caffeine from coffee.[17] Carbon dioxide has attracted attention in the pharmaceutical and other chemical processing industries as a less toxic alternative to more traditional solvents such as organochlorides. It is also used by some dry cleaners for this reason. It is used in the preparation of some aerogels because of the properties of supercritical carbon dioxide.

Medical and pharmacological uses

In medicine, up to 5% carbon dioxide (130 times atmospheric concentration) is added to oxygen for stimulation of breathing after apnea and to stabilize the O
2
/CO
2
balance in blood.

Carbon dioxide can be mixed with up to 50% oxygen, forming an inhalable gas; this is known as Carbogen and has a variety of medical and research uses.

Another medical use are the mofette, dry spas that use carbon dioxide from post-volcanic discharge for therapeutic purposes.

Energy

Supercritical CO2 is used as the working fluid in the Allam power cycle engine.

Fossil fuel recovery

Carbon dioxide is used in enhanced oil recovery where it is injected into or adjacent to producing oil wells, usually under supercritical conditions, when it becomes miscible with the oil. This approach can increase original oil recovery by reducing residual oil saturation by between 7% to 23% additional to primary extraction.[67] It acts as both a pressurizing agent and, when dissolved into the underground crude oil, significantly reduces its viscosity, and changing surface chemistry enabling the oil to flow more rapidly through the reservoir to the removal well.[68] In mature oil fields, extensive pipe networks are used to carry the carbon dioxide to the injection points.

In enhanced coal bed methane recovery, carbon dioxide would be pumped into the coal seam to displace methane, as opposed to current methods which primarily rely on the removal of water (to reduce pressure) to make the coal seam release its trapped methane.[69]

Bio transformation into fuel

It has been proposed that CO2 from power generation be bubbled into ponds to stimulate growth of algae that could then be converted into biodiesel fuel.[70] A strain of the cyanobacterium Synechococcus elongatus has been genetically engineered to produce the fuels isobutyraldehyde and isobutanol from CO2 using photosynthesis.[71]

Researchers have developed a process called electrolysis, using enzymes isolated from bacteria to power the chemical reactions which convert CO2 into fuels.[72][73][74]

Refrigerant

What elements make up carbon dioxide

Comparison of the pressure–temperature phase diagrams of carbon dioxide (red) and water (blue) as a log-lin chart with phase transitions points at 1 atmosphere

Liquid and solid carbon dioxide are important refrigerants, especially in the food industry, where they are employed during the transportation and storage of ice cream and other frozen foods. Solid carbon dioxide is called "dry ice" and is used for small shipments where refrigeration equipment is not practical. Solid carbon dioxide is always below −78.5 °C (−109.3 °F) at regular atmospheric pressure, regardless of the air temperature.

Liquid carbon dioxide (industry nomenclature R744 or R-744) was used as a refrigerant prior to the use[citation needed] of dichlorodifluoromethane (R12, a chlorofluorocarbon (CFC) compound). CO2 might enjoy a renaissance because one of the main substitutes to CFCs, 1,1,1,2-tetrafluoroethane (R134a, a hydrofluorocarbon (HFC) compound) contributes to climate change more than CO2 does. CO2 physical properties are highly favorable for cooling, refrigeration, and heating purposes, having a high volumetric cooling capacity. Due to the need to operate at pressures of up to 130 bars (1,900 psi; 13,000 kPa), CO2 systems require highly mechanically resistant reservoirs and components that have already been developed for mass production in many sectors. In automobile air conditioning, in more than 90% of all driving conditions for latitudes higher than 50°, CO2 (R744) operates more efficiently than systems using HFCs (e.g., R134a). Its environmental advantages (GWP of 1, non-ozone depleting, non-toxic, non-flammable) could make it the future working fluid to replace current HFCs in cars, supermarkets, and heat pump water heaters, among others. Coca-Cola has fielded CO2-based beverage coolers and the U.S. Army is interested in CO2 refrigeration and heating technology.[75][76]

Minor uses

What elements make up carbon dioxide

Carbon dioxide is the lasing medium in a carbon-dioxide laser, which is one of the earliest type of lasers.

Carbon dioxide can be used as a means of controlling the pH of swimming pools,[77] by continuously adding gas to the water, thus keeping the pH from rising. Among the advantages of this is the avoidance of handling (more hazardous) acids. Similarly, it is also used in the maintaining reef aquaria, where it is commonly used in calcium reactors to temporarily lower the pH of water being passed over calcium carbonate in order to allow the calcium carbonate to dissolve into the water more freely, where it is used by some corals to build their skeleton.

Used as the primary coolant in the British advanced gas-cooled reactor for nuclear power generation.

Carbon dioxide induction is commonly used for the euthanasia of laboratory research animals. Methods to administer CO2 include placing animals directly into a closed, prefilled chamber containing CO2, or exposure to a gradually increasing concentration of CO2. The American Veterinary Medical Association's 2020 guidelines for carbon dioxide induction state that a displacement rate of 30% to 70% of the chamber or cage volume per minute is optimal for the humane euthanasia of small rodents.[78]: 5, 31 Percentages of CO2 vary for different species, based on identified optimal percentages to minimize distress.[78]: 22

Carbon dioxide is also used in several related cleaning and surface-preparation techniques.

In Earth's atmosphere

What elements make up carbon dioxide

Carbon dioxide in Earth's atmosphere is a trace gas, having a global average concentration of 415 parts per million by volume (or 630 parts per million by mass) as of the end of year 2020.[80][81] Atmospheric CO2 concentrations fluctuate slightly with the seasons, falling during the Northern Hemisphere spring and summer as plants consume the gas and rising during northern autumn and winter as plants go dormant or die and decay. Concentrations also vary on a regional basis, most strongly near the ground with much smaller variations aloft. In urban areas concentrations are generally higher[82] and indoors they can reach 10 times background levels. CO2 emissions have also led to the stratosphere contracting by 400 meters since 1980, which could affect satellite operations, GPS systems and radio communications.[83]

The concentration of carbon dioxide has risen due to human activities.[84] The extraction and burning of fossil fuels, using carbon that has been sequestered for many millions of years in the lithosphere, has caused the atmospheric concentration of CO2 to increase by about 50% since the beginning of the age of industrialization up to year 2020.[85][86] Most CO2 from human activities is released from burning coal, petroleum, and natural gas. Other large anthropogenic sources include cement production, deforestation, and biomass burning. Human activities emit over 30 billion tons of CO2 (9 billion tons of fossil carbon) per year, while volcanoes emit only between 0.2 and 0.3 billion tons of CO2.[87][88] Human activities have caused CO2 to increase above levels not seen in hundreds of thousands of years. Currently, about half of the carbon dioxide released from the burning of fossil fuels remains in the atmosphere and is not absorbed by vegetation and the oceans.[89][90][91][92]

While transparent to visible light, carbon dioxide is a greenhouse gas, absorbing and emitting infrared radiation at its two infrared-active vibrational frequencies (see the section "Structure and bonding" above). Light emission from the Earth's surface is most intense in the infrared region between 200 and 2500 cm−1,[93] as opposed to light emission from the much hotter Sun which is most intense in the visible region. Absorption of infrared light at the vibrational frequencies of atmospheric CO2 traps energy near the surface, warming the surface and the lower atmosphere. Less energy reaches the upper atmosphere, which is therefore cooler because of this absorption.[94]

What elements make up carbon dioxide

Annual CO2 flows from anthropogenic sources (left) into Earth's atmosphere, land, and ocean sinks (right) since the 1960s. Units in equivalent gigatonnes carbon per year.[86]

Increases in atmospheric concentrations of CO2 and other long-lived greenhouse gases such as methane, nitrous oxide and ozone have strengthened their absorption and emission of infrared radiation, causing the rise in average global temperature since the mid-20th century. Carbon dioxide is of greatest concern because it exerts a larger overall warming influence than all of these other gases combined.[85] It furthermore has an atmospheric lifetime that increases with the cumulative amount of fossil carbon extracted and burned, due to the imbalance that this activity has imposed on Earth's fast carbon cycle.[95] This means that some fraction (a projected 20-35%) of the fossil carbon transferred thus far will persist in the atmosphere as elevated CO2 levels for many thousands of years after these carbon transfer activities begin to subside.[96][97][98] Not only do increasing CO2 concentrations lead to increases in global surface temperature, but increasing global temperatures also cause increasing concentrations of carbon dioxide. This produces a positive feedback for changes induced by other processes such as orbital cycles.[99] Five hundred million years ago the CO2 concentration was 20 times greater than today, decreasing to 4–5 times during the Jurassic period and then slowly declining with a particularly swift reduction occurring 49 million years ago.[100][101]

Local concentrations of carbon dioxide can reach high values near strong sources, especially those that are isolated by surrounding terrain. At the Bossoleto hot spring near Rapolano Terme in Tuscany, Italy, situated in a bowl-shaped depression about 100 m (330 ft) in diameter, concentrations of CO2 rise to above 75% overnight, sufficient to kill insects and small animals. After sunrise the gas is dispersed by convection.[102] High concentrations of CO2 produced by disturbance of deep lake water saturated with CO2 are thought to have caused 37 fatalities at Lake Monoun, Cameroon in 1984 and 1700 casualties at Lake Nyos, Cameroon in 1986.[103]

In the oceans

What elements make up carbon dioxide

Pterapod shell dissolved in seawater adjusted to an ocean chemistry projected for the year 2100.

Carbon dioxide dissolves in the ocean to form carbonic acid (H2CO3), bicarbonate (HCO3−) and carbonate (CO32−). There is about fifty times as much carbon dioxide dissolved in the oceans as exists in the atmosphere. The oceans act as an enormous carbon sink, and have taken up about a third of CO2 emitted by human activity.[104]

As the concentration of carbon dioxide increases in the atmosphere, the increased uptake of carbon dioxide into the oceans is causing a measurable decrease in the pH of the oceans, which is referred to as ocean acidification. This reduction in pH affects biological systems in the oceans, primarily oceanic calcifying organisms. These effects span the food chain from autotrophs to heterotrophs and include organisms such as coccolithophores, corals, foraminifera, echinoderms, crustaceans and mollusks. Under normal conditions, calcium carbonate is stable in surface waters since the carbonate ion is at supersaturating concentrations. However, as ocean pH falls, so does the concentration of this ion, and when carbonate becomes undersaturated, structures made of calcium carbonate are vulnerable to dissolution.[105] Corals,[106][107][108] coccolithophore algae,[109][110][111][112] coralline algae,[113] foraminifera,[114] shellfish[115] and pteropods[116] experience reduced calcification or enhanced dissolution when exposed to elevated CO
2
.

Gas solubility decreases as the temperature of water increases (except when both pressure exceeds 300 bar and temperature exceeds 393 K, only found near deep geothermal vents)[117] and therefore the rate of uptake from the atmosphere decreases as ocean temperatures rise.

Most of the CO2 taken up by the ocean, which is about 30% of the total released into the atmosphere,[118] forms carbonic acid in equilibrium with bicarbonate. Some of these chemical species are consumed by photosynthetic organisms that remove carbon from the cycle. Increased CO2 in the atmosphere has led to decreasing alkalinity of seawater, and there is concern that this may adversely affect organisms living in the water. In particular, with decreasing alkalinity, the availability of carbonates for forming shells decreases,[119] although there's evidence of increased shell production by certain species under increased CO2 content.[120]

The U.S. National Oceanic and Atmospheric Administration (NOAA) states in their May 2008 "State of the science fact sheet for ocean acidification"[121] that:

The oceans have absorbed about 50% of the carbon dioxide (CO2) released from the burning of fossil fuels, resulting in chemical reactions that lower ocean pH. This has caused an increase in hydrogen ion (acidity) of about 30% since the start of the industrial age through a process known as "ocean acidification". A growing number of studies have demonstrated adverse impacts on marine organisms, including:

  • The rate at which reef-building corals produce their skeletons decreases, while production of numerous varieties of jellyfish increases.
  • The ability of marine algae and free-swimming zooplankton to maintain protective shells is reduced.
  • The survival of larval marine species, including commercial fish and shellfish, is reduced.

Also, the Intergovernmental Panel on Climate Change (IPCC) writes in their Climate Change 2007: Synthesis Report:[122]

The uptake of anthropogenic carbon since 1750 has led to the ocean becoming more acidic with an average decrease in pH of 0.1 units. Increasing atmospheric CO2 concentrations lead to further acidification ... While the effects of observed ocean acidification on the marine biosphere are as yet undocumented, the progressive acidification of oceans is expected to have negative impacts on marine shell-forming organisms (e.g. corals) and their dependent species.

Some marine calcifying organisms (including coral reefs) have been singled out by major research agencies, including NOAA, the OSPAR Commission, the Northwest Association of Networked Ocean Observing Systems, and the IPCC, because their most current research shows that ocean acidification should be expected to impact them negatively.[123]

Carbon dioxide is also introduced into the oceans through hydrothermal vents. The Champagne hydrothermal vent, found at the Northwest Eifuku volcano in the Mariana Trench, produces almost pure liquid carbon dioxide, one of only two known sites in the world as of 2004, the other being in the Okinawa Trough.[124] The finding of a submarine lake of liquid carbon dioxide in the Okinawa Trough was reported in 2006.[125]

Biological role

Carbon dioxide is an end product of cellular respiration in organisms that obtain energy by breaking down sugars, fats and amino acids with oxygen as part of their metabolism. This includes all plants, algae and animals and aerobic fungi and bacteria. In vertebrates, the carbon dioxide travels in the blood from the body's tissues to the skin (e.g., amphibians) or the gills (e.g., fish), from where it dissolves in the water, or to the lungs from where it is exhaled. During active photosynthesis, plants can absorb more carbon dioxide from the atmosphere than they release in respiration.

Photosynthesis and carbon fixation

What elements make up carbon dioxide

Carbon fixation is a biochemical process by which atmospheric carbon dioxide is incorporated by plants, algae and (cyanobacteria) into energy-rich organic molecules such as glucose, thus creating their own food by photosynthesis. Photosynthesis uses carbon dioxide and water to produce sugars from which other organic compounds can be constructed, and oxygen is produced as a by-product.

Ribulose-1,5-bisphosphate carboxylase oxygenase, commonly abbreviated to RuBisCO, is the enzyme involved in the first major step of carbon fixation, the production of two molecules of 3-phosphoglycerate from CO2 and ribulose bisphosphate, as shown in the diagram at left.

RuBisCO is thought to be the single most abundant protein on Earth.[126]

Phototrophs use the products of their photosynthesis as internal food sources and as raw material for the biosynthesis of more complex organic molecules, such as polysaccharides, nucleic acids and proteins. These are used for their own growth, and also as the basis of the food chains and webs that feed other organisms, including animals such as ourselves. Some important phototrophs, the coccolithophores synthesise hard calcium carbonate scales.[127] A globally significant species of coccolithophore is Emiliania huxleyi whose calcite scales have formed the basis of many sedimentary rocks such as limestone, where what was previously atmospheric carbon can remain fixed for geological timescales.

What elements make up carbon dioxide

Overview of photosynthesis and respiration. Carbon dioxide (at right), together with water, form oxygen and organic compounds (at left) by photosynthesis, which can be respired to water and (CO2).

Plants can grow as much as 50 percent faster in concentrations of 1,000 ppm CO2 when compared with ambient conditions, though this assumes no change in climate and no limitation on other nutrients.[128] Elevated CO2 levels cause increased growth reflected in the harvestable yield of crops, with wheat, rice and soybean all showing increases in yield of 12–14% under elevated CO2 in FACE experiments.[129][130]

Increased atmospheric CO2 concentrations result in fewer stomata developing on plants[131] which leads to reduced water usage and increased water-use efficiency.[132] Studies using FACE have shown that CO2 enrichment leads to decreased concentrations of micronutrients in crop plants.[133] This may have knock-on effects on other parts of ecosystems as herbivores will need to eat more food to gain the same amount of protein.[134]

The concentration of secondary metabolites such as phenylpropanoids and flavonoids can also be altered in plants exposed to high concentrations of CO2.[135][136]

Plants also emit CO2 during respiration, and so the majority of plants and algae, which use C3 photosynthesis, are only net absorbers during the day. Though a growing forest will absorb many tons of CO2 each year, a mature forest will produce as much CO2 from respiration and decomposition of dead specimens (e.g., fallen branches) as is used in photosynthesis in growing plants.[137] Contrary to the long-standing view that they are carbon neutral, mature forests can continue to accumulate carbon[138] and remain valuable carbon sinks, helping to maintain the carbon balance of Earth's atmosphere. Additionally, and crucially to life on earth, photosynthesis by phytoplankton consumes dissolved CO2 in the upper ocean and thereby promotes the absorption of CO2 from the atmosphere.[139]

Toxicity

What elements make up carbon dioxide

Carbon dioxide content in fresh air (averaged between sea-level and 10 kPa level, i.e., about 30 km (19 mi) altitude) varies between 0.036% (360 ppm) and 0.041% (412 ppm), depending on the location.[141][clarification needed]

CO2 is an asphyxiant gas and not classified as toxic or harmful in accordance with Globally Harmonized System of Classification and Labelling of Chemicals standards of United Nations Economic Commission for Europe by using the OECD Guidelines for the Testing of Chemicals. In concentrations up to 1% (10,000 ppm), it will make some people feel drowsy and give the lungs a stuffy feeling.[140] Concentrations of 7% to 10% (70,000 to 100,000 ppm) may cause suffocation, even in the presence of sufficient oxygen, manifesting as dizziness, headache, visual and hearing dysfunction, and unconsciousness within a few minutes to an hour.[142] The physiological effects of acute carbon dioxide exposure are grouped together under the term hypercapnia, a subset of asphyxiation.

Because it is heavier than air, in locations where the gas seeps from the ground (due to sub-surface volcanic or geothermal activity) in relatively high concentrations, without the dispersing effects of wind, it can collect in sheltered/pocketed locations below average ground level, causing animals located therein to be suffocated. Carrion feeders attracted to the carcasses are then also killed. Children have been killed in the same way near the city of Goma by CO2 emissions from the nearby volcano Mount Nyiragongo.[143] The Swahili term for this phenomenon is mazuku.

What elements make up carbon dioxide

Adaptation to increased concentrations of CO2 occurs in humans, including modified breathing and kidney bicarbonate production, in order to balance the effects of blood acidification (acidosis). Several studies suggested that 2.0 percent inspired concentrations could be used for closed air spaces (e.g. a submarine) since the adaptation is physiological and reversible, as deterioration in performance or in normal physical activity does not happen at this level of exposure for five days.[144][145] Yet, other studies show a decrease in cognitive function even at much lower levels.[146][147] Also, with ongoing respiratory acidosis, adaptation or compensatory mechanisms will be unable to reverse such condition.

Below 1%

There are few studies of the health effects of long-term continuous CO2 exposure on humans and animals at levels below 1%. Occupational CO2 exposure limits have been set in the United States at 0.5% (5000 ppm) for an eight-hour period.[148] At this CO2 concentration, International Space Station crew experienced headaches, lethargy, mental slowness, emotional irritation, and sleep disruption.[149] Studies in animals at 0.5% CO2 have demonstrated kidney calcification and bone loss after eight weeks of exposure.[150] A study of humans exposed in 2.5 hour sessions demonstrated significant negative effects on cognitive abilities at concentrations as low as 0.1% (1000 ppm) CO2 likely due to CO2 induced increases in cerebral blood flow.[146] Another study observed a decline in basic activity level and information usage at 1000 ppm, when compared to 500 ppm.[147] However a review of the literature found that most studies on the phenomenon of carbon dioxide induced cognitive impairment to have a small effect on high-level decision making and most of the studies were confounded by inadequate study designs, environmental comfort, uncertainties in exposure doses and differing cognitive assessments used.[151] Similarly a study on the effects of the concentration of CO2 in motorcycle helmets has been criticized for having dubious methodology in not noting the self-reports of motorcycle riders and taking measurements using mannequins. Further when normal motorcycle conditions were achieved (such as highway or city speeds) or the visor was raised the concentration of CO2 declined to safe levels (0.2%).[152][153]

Ventilation

What elements make up carbon dioxide

Poor ventilation is one of the main causes of excessive CO2 concentrations in closed spaces. Carbon dioxide differential above outdoor concentrations at steady state conditions (when the occupancy and ventilation system operation are sufficiently long that CO2 concentration has stabilized) are sometimes used to estimate ventilation rates per person.[citation needed] Higher CO2 concentrations are associated with occupant health, comfort and performance degradation.[154][155] ASHRAE Standard 62.1–2007 ventilation rates may result in indoor concentrations up to 2,100 ppm above ambient outdoor conditions. Thus if the outdoor concentration is 400 ppm, indoor concentrations may reach 2,500 ppm with ventilation rates that meet this industry consensus standard. Concentrations in poorly ventilated spaces can be found even higher than this (range of 3,000 or 4,000 ppm).

Miners, who are particularly vulnerable to gas exposure due to insufficient ventilation, referred to mixtures of carbon dioxide and nitrogen as "blackdamp," "choke damp" or "stythe." Before more effective technologies were developed, miners would frequently monitor for dangerous levels of blackdamp and other gases in mine shafts by bringing a caged canary with them as they worked. The canary is more sensitive to asphyxiant gases than humans, and as it became unconscious would stop singing and fall off its perch. The Davy lamp could also detect high levels of blackdamp (which sinks, and collects near the floor) by burning less brightly, while methane, another suffocating gas and explosion risk, would make the lamp burn more brightly.

In February 2020, three people died from suffocation at a party in Moscow when dry ice (frozen CO2) was added to a swimming pool to cool it down.[156] A similar accident occurred in 2018 when a woman died from CO2 fumes emanating from the large amount of dry ice she was transporting in her car.[157]

Human physiology

Content

Reference ranges or averages for partial pressures of carbon dioxide (abbreviated pCO2)
Blood compartment(kPa)(mm Hg)
Venous blood carbon dioxide 5.5–6.8 41–51[158]
Alveolar pulmonary
gas pressures
4.8 36
Arterial blood carbon dioxide4.7–6.0 35–45[158]

The body produces approximately 2.3 pounds (1.0 kg) of carbon dioxide per day per person,[159] containing 0.63 pounds (290 g) of carbon. In humans, this carbon dioxide is carried through the venous system and is breathed out through the lungs, resulting in lower concentrations in the arteries. The carbon dioxide content of the blood is often given as the partial pressure, which is the pressure which carbon dioxide would have had if it alone occupied the volume.[160] In humans, the blood carbon dioxide contents is shown in the adjacent table.

Transport in the blood

CO2 is carried in blood in three different ways. (Exact percentages vary between arterial and venous blood).

  • Majority (about 70% to 80%) is converted to bicarbonate ions HCO
    3
    by the enzyme carbonic anhydrase in the red blood cells,[161] by the reaction CO2 + H
    2
    O
    H
    2
    CO
    3
    H+
    + HCO
    3
    .
  • 5–10% is dissolved in blood plasma[161]
  • 5–10% is bound to hemoglobin as carbamino compounds[161]

Hemoglobin, the main oxygen-carrying molecule in red blood cells, carries both oxygen and carbon dioxide. However, the CO2 bound to hemoglobin does not bind to the same site as oxygen. Instead, it combines with the N-terminal groups on the four globin chains. However, because of allosteric effects on the hemoglobin molecule, the binding of CO2 decreases the amount of oxygen that is bound for a given partial pressure of oxygen. This is known as the Haldane Effect, and is important in the transport of carbon dioxide from the tissues to the lungs. Conversely, a rise in the partial pressure of CO2 or a lower pH will cause offloading of oxygen from hemoglobin, which is known as the Bohr effect.

Regulation of respiration

Carbon dioxide is one of the mediators of local autoregulation of blood supply. If its concentration is high, the capillaries expand to allow a greater blood flow to that tissue.[162]

Bicarbonate ions are crucial for regulating blood pH. A person's breathing rate influences the level of CO2 in their blood. Breathing that is too slow or shallow causes respiratory acidosis, while breathing that is too rapid leads to hyperventilation, which can cause respiratory alkalosis.[163]

Although the body requires oxygen for metabolism, low oxygen levels normally do not stimulate breathing. Rather, breathing is stimulated by higher carbon dioxide levels. As a result, breathing low-pressure air or a gas mixture with no oxygen at all (such as pure nitrogen) can lead to loss of consciousness without ever experiencing air hunger. This is especially perilous for high-altitude fighter pilots. It is also why flight attendants instruct passengers, in case of loss of cabin pressure, to apply the oxygen mask to themselves first before helping others; otherwise, one risks losing consciousness.[161]

The respiratory centers try to maintain an arterial CO2 pressure of 40 mm Hg. With intentional hyperventilation, the CO2 content of arterial blood may be lowered to 10–20 mm Hg (the oxygen content of the blood is little affected), and the respiratory drive is diminished. This is why one can hold one's breath longer after hyperventilating than without hyperventilating. This carries the risk that unconsciousness may result before the need to breathe becomes overwhelming, which is why hyperventilation is particularly dangerous before free diving.[164]

See also

  • Arterial blood gas
  • Azolla event – Hypothetical geoclimatic event, 49 M yrs ago
  • Bosch reaction
  • Carbon dioxide removal – Removal of atmospheric carbon dioxide (from the atmosphere)
  • Carbon dioxide sensor
  • Carbon sequestration – Capture and long-term storage of atmospheric carbon dioxide
  • Cave of Dogs – Cave near Naples, Italy
  • Emission standards
  • Indoor air quality – Air quality within and around buildings and structures
  • Kaya identity – Identity regarding anthropogenic carbon dioxide emissions
  • Lake Kivu – Meromictic lake in the East African Rift valley
  • List of least carbon efficient power stations
  • List of countries by carbon dioxide emissions
  • Meromictic lake – Permanently stratified lake with layers of water that do not intermix
  • pCO2 – Partial pressure of carbon dioxide, often used in reference to blood
  • Gilbert Plass – Canadian physicist (early work on CO2 and climate change)
  • Sabatier reaction – Methanation process of carbon dioxide with hydrogen
  • NASA's Orbiting Carbon Observatory 2
  • Greenhouse Gases Observing Satellite – Earth observation satellite
  • Soil gas

References

  1. ^ a b c "Carbon Dioxide" (PDF). Air Products. Archived from the original (PDF) on 29 July 2020. Retrieved 28 April 2017.
  2. ^ a b c d e f g h i Span R, Wagner W (1 November 1996). "A New Equation of State for Carbon Dioxide Covering the Fluid Region from the Triple‐Point Temperature to 1100 K at Pressures up to 800 MPa". Journal of Physical and Chemical Reference Data. 25 (6): 1519. Bibcode:1996JPCRD..25.1509S. doi:10.1063/1.555991.
  3. ^ Touloukian YS, Liley PE, Saxena SC (1970). "Thermophysical properties of matter - the TPRC data series". Thermal Conductivity - Nonmetallic Liquids and Gases. Data book. 3.
  4. ^ Schäfer M, Richter M, Span R (2015). "Measurements of the viscosity of carbon dioxide at temperatures from (253.15 to 473.15) K with pressures up to 1.2 MPa". The Journal of Chemical Thermodynamics. 89: 7–15. doi:10.1016/j.jct.2015.04.015.
  5. ^ a b c NIOSH Pocket Guide to Chemical Hazards. "#0103". National Institute for Occupational Safety and Health (NIOSH).
  6. ^ "Carbon dioxide". Immediately Dangerous to Life or Health Concentrations (IDLH). National Institute for Occupational Safety and Health (NIOSH).
  7. ^ "Safety Data Sheet – Carbon Dioxide Gas – version 0.03 11/11" (PDF). AirGas.com. 12 February 2018. Archived (PDF) from the original on 4 August 2018. Retrieved 4 August 2018.
  8. ^ "Carbon dioxide, refrigerated liquid" (PDF). Praxair. p. 9. Archived from the original (PDF) on 29 July 2018. Retrieved 26 July 2018.
  9. ^ Eggleton T (2013). A Short Introduction to Climate Change. Cambridge University Press. p. 52. ISBN 9781107618763. Archived from the original on 23 July 2021. Retrieved 9 November 2020.
  10. ^ "Carbon Dioxide Concentration". Climate Change: Vital Signs of the Planet. NASA. Archived from the original on 23 June 2021. Retrieved 23 June 2021.
  11. ^ IPCC AR6 WG3 Summary for Policymakers 2022 (PDF). Intergovernmental Panel on Climate Change. 2022. Figure SPM.1. Archived (PDF) from the original on 10 October 2022.
  12. ^ Ocean Acidification: A National Strategy to Meet the Challenges of a Changing Ocean. Washington, DC: National Academies Press. 22 April 2010. doi:10.17226/12904. ISBN 978-0-309-15359-1. Archived from the original on 5 February 2016. Retrieved 29 February 2016.
  13. ^ Kaufman DG, Franz CM (1996). Biosphere 2000: protecting our global environment. Kendall/Hunt Pub. Co. ISBN 978-0-7872-0460-0.
  14. ^ "Food Factories". www.legacyproject.org. Archived from the original on 12 August 2017. Retrieved 10 October 2011.
  15. ^ IPCC (2021). "Summary for Policymakers" (PDF). Climate Change 2021: The Physical Science Basis. p. 20. Archived (PDF) from the original on 10 October 2022.
  16. ^ Myles, Allen (September 2020). "The Oxford Principles for Net Zero Aligned Carbon Offsetting" (PDF). Archived (PDF) from the original on 2 October 2020. Retrieved 10 December 2021.
  17. ^ a b Tsotsas E, Mujumdar AS (2011). Modern drying technology. Vol. 3: Product quality and formulation. John Wiley & Sons. ISBN 978-3-527-31558-1. Archived from the original on 21 March 2020. Retrieved 3 December 2019.
  18. ^ Mikhail M, Wang B, Jalain R, Cavadias S, Tatoulian M, Ognier S, Gálvez ME, Da Costa P (1 April 2019). "Plasma-catalytic hybrid process for CO2 methanation: optimization of operation parameters". Reaction Kinetics, Mechanisms and Catalysis. 126 (2): 629–643. doi:10.1007/s11144-018-1508-8. S2CID 104301429.
  19. ^ "Catalysts for climate protection". Fraunhofer Institute for Interfacial Engineering and Biotechnology. 19 August 2019. Archived from the original on 1 October 2021. Retrieved 19 October 2019.
  20. ^ Voiry D, Shin HS, Loh KP, Chhowalla M (2018). "Low-dimensional catalysts for hydrogen evolution and CO2 reduction". Nature Reviews Chemistry. 2 (1): 0105. doi:10.1038/s41570-017-0105.
  21. ^ Gomez E, Yan B, Kattel S, Chen JG (10 September 2019). "Carbon dioxide reduction in tandem with light-alkane dehydrogenation". Nature Reviews Chemistry. 3 (11): 638–649. doi:10.1038/s41570-019-0128-9. OSTI 1580234. S2CID 202159972. Archived from the original on 15 March 2020. Retrieved 19 October 2019.
  22. ^ Csepei LI, Muhler M (2011). Kinetic studies of propane oxidation on Mo and V based mixed oxide catalysts (PDF) (PhD thesis). Technical University of Berlin. Archived (PDF) from the original on 30 May 2016. Retrieved 9 July 2017.
  23. ^ Amakawa K, Kolen'ko YV, Villa A, Schuster ME, Csepei LI, Weinberg G, et al. (2013). "Multifunctionality of Crystalline MoV(TeNb) M1 Oxide Catalysts in Selective Oxidation of Propane and Benzyl Alcohol". ACS Catalysis. 3 (6): 1103–1113. doi:10.1021/cs400010q. Archived from the original on 22 October 2018. Retrieved 9 July 2017.
  24. ^ d'Alnoncourt RN, Csepei LI, Hävecker M, Girgsdies F, Schuster ME, Schlögl R, Trunschke A (2014). "The reaction network in propane oxidation over phase-pure MoVTeNb M1 oxide catalysts" (PDF). Journal of Catalysis. 311: 369–385. doi:10.1016/j.jcat.2013.12.008. hdl:11858/00-001M-0000-0014-F434-5. Archived from the original (PDF) on 15 February 2016. Retrieved 9 July 2017.
  25. ^ Spritzler F (3 November 2019). "Carbonated (Sparkling) Water: Good or Bad?". healthline.com. Archived from the original on 10 May 2020.
  26. ^ Harris D (September 1910). "The Pioneer in the Hygiene of Ventilation". The Lancet. 176 (4542): 906–908. doi:10.1016/S0140-6736(00)52420-9. Archived from the original on 17 March 2020. Retrieved 6 December 2019.
  27. ^ Almqvist E (2003). History of industrial gases. Springer. p. 93. ISBN 978-0-306-47277-0.
  28. ^ Priestley J, Hey W (1772). "Observations on Different Kinds of Air". Philosophical Transactions. 62: 147–264. doi:10.1098/rstl.1772.0021. S2CID 186210131. Archived from the original on 7 June 2010. Retrieved 11 October 2007.
  29. ^ Davy H (1823). "On the Application of Liquids Formed by the Condensation of Gases as Mechanical Agents". Philosophical Transactions. 113: 199–205. doi:10.1098/rstl.1823.0020. JSTOR 107649.
  30. ^ Thilorier AJ (1835). "Solidification de l'Acide carbonique". Comptes Rendus. 1: 194–196. Archived from the original on 2 September 2017. Retrieved 1 September 2017.
  31. ^ Thilorier AJ (1836). "Solidification of carbonic acid". The London and Edinburgh Philosophical Magazine. 8 (48): 446–447. doi:10.1080/14786443608648911. Archived from the original on 2 May 2016. Retrieved 15 November 2015.
  32. ^ a b Greenwood NN, Earnshaw A (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. pp. 305–314. ISBN 978-0-08-037941-8.
  33. ^ Atkins P, de Paula J (2006). Physical Chemistry (8th ed.). W.H. Freeman. pp. 461, 464. ISBN 978-0-7167-8759-4.
  34. ^ Siegmann B, Werner U, Lutz HO, Mann R (2002). "Complete Coulomb fragmentation of CO2 in collisions with 5.9 MeV u−1 Xe18+ and Xe43+". J Phys B Atom Mol Opt Phys. 35 (17): 3755. Bibcode:2002JPhB...35.3755S. doi:10.1088/0953-4075/35/17/311. S2CID 250782825.
  35. ^ Jensen P, Spanner M, Bunker PR (2020). "The CO2 molecule is never linear−". J Mol Struct. 1212: 128087. Bibcode:2020JMoSt121228087J. doi:10.1016/j.molstruc.2020.128087.
  36. ^ Jolly WL (1984). Modern Inorganic Chemistry. McGraw-Hill. p. 196. ISBN 978-0-07-032760-3.
  37. ^ Li Z, Mayer RJ, Ofial AR, Mayr H (May 2020). "From Carbodiimides to Carbon Dioxide: Quantification of the Electrophilic Reactivities of Heteroallenes". Journal of the American Chemical Society. 142 (18): 8383–8402. doi:10.1021/jacs.0c01960. PMID 32338511. S2CID 216557447.
  38. ^ Aresta M, ed. (2010). Carbon Dioxide as a Chemical Feedstock. Weinheim: Wiley-VCH. ISBN 978-3-527-32475-0.
  39. ^ Finn C, Schnittger S, Yellowlees LJ, Love JB (February 2012). "Molecular approaches to the electrochemical reduction of carbon dioxide" (PDF). Chemical Communications. 48 (10): 1392–1399. doi:10.1039/c1cc15393e. hdl:20.500.11820/b530915d-451c-493c-8251-da2ea2f50912. PMID 22116300. S2CID 14356014. Archived (PDF) from the original on 19 April 2021. Retrieved 6 December 2019.
  40. ^ "Gases – Densities". Engineering Toolbox. Archived from the original on 2 March 2006. Retrieved 21 November 2020.
  41. ^ Santoro M, Gorelli FA, Bini R, Ruocco G, Scandolo S, Crichton WA (June 2006). "Amorphous silica-like carbon dioxide". Nature. 441 (7095): 857–860. Bibcode:2006Natur.441..857S. doi:10.1038/nature04879. PMID 16778885. S2CID 4363092.
  42. ^ a b Pierantozzi R (2001). "Carbon Dioxide". Kirk-Othmer Encyclopedia of Chemical Technology. Wiley. doi:10.1002/0471238961.0301180216090518.a01.pub2. ISBN 978-0-471-23896-6.
  43. ^ Strassburger J (1969). Blast Furnace Theory and Practice. New York: American Institute of Mining, Metallurgical, and Petroleum Engineers. ISBN 978-0-677-10420-1.
  44. ^ Topham S (2000). "Carbon Dioxide". Ullmann's Encyclopedia of Industrial Chemistry. doi:10.1002/14356007.a05_165. ISBN 3527306730.
  45. ^ "CO2 shortage: Food industry calls for government action". BBC. 21 June 2018. Archived from the original on 23 May 2021. Retrieved 24 June 2018.
  46. ^ "Collecting and using biogas from landfills". U.S. Energy Information Administration. 11 January 2017. Archived from the original on 11 July 2018. Retrieved 22 November 2015.
  47. ^ "Facts About Landfill Gas" (PDF). U.S. Environmental Protection Agency. January 2000. Archived (PDF) from the original on 23 September 2015. Retrieved 4 September 2015.
  48. ^ "IPCC Special Report on Carbon dioxide Capture and Storage" (PDF). The Intergovernmental Panel on Climate Change. Archived from the original (PDF) on 24 September 2015. Retrieved 4 September 2015.
  49. ^ Morrison RT, Boyd RN (1983). Organic Chemistry (4th ed.). Allyn and Bacon. pp. 976–977. ISBN 978-0-205-05838-9.
  50. ^ Badwal SP, Giddey SS, Munnings C, Bhatt AI, Hollenkamp AF (24 September 2014). "Emerging electrochemical energy conversion and storage technologies". Frontiers in Chemistry. 2: 79. Bibcode:2014FrCh....2...79B. doi:10.3389/fchem.2014.00079. PMC 4174133. PMID 25309898.
  51. ^ Whiting D, Roll M, Vickerman L (August 2010). "Plant Growth Factors: Photosynthesis, Respiration, and Transpiration". CMG GardenNotes. Colorado Master Gardener Program. Archived from the original on 2 September 2014. Retrieved 10 October 2011.
  52. ^ Waggoner PE (February 1994). "Carbon dioxide". How Much Land Can Ten Billion People Spare for Nature?. Archived from the original on 12 October 2011. Retrieved 10 October 2011.
  53. ^ Stafford N (August 2007). "Future crops: the other greenhouse effect". Nature. 448 (7153): 526–528. Bibcode:2007Natur.448..526S. doi:10.1038/448526a. PMID 17671477. S2CID 9845813.
  54. ^ UK Food Standards Agency: "Current EU approved additives and their E Numbers". Archived from the original on 7 October 2010. Retrieved 27 October 2011.
  55. ^ US Food and Drug Administration: "Food Additive Status List". Food and Drug Administration. Archived from the original on 4 November 2017. Retrieved 13 June 2015.
  56. ^ Australia New Zealand Food Standards Code"Standard 1.2.4 – Labelling of ingredients". Archived from the original on 19 January 2012. Retrieved 27 October 2011.
  57. ^ Futurific Leading Indicators Magazine. Vol. 1. CRAES LLC. ISBN 978-0-9847670-1-4. Archived from the original on 15 August 2021. Retrieved 9 November 2020.
  58. ^ Vijay GP (25 September 2015). Indian Breads: A Comprehensive Guide to Traditional and Innovative Indian Breads. Westland. ISBN 978-93-85724-46-6.
  59. ^ "Scientists Discover Protein Receptor For Carbonation Taste". ScienceDaily. 16 October 2009. Archived from the original on 29 March 2020. Retrieved 29 March 2020.
  60. ^ Coghlan A (3 February 2018). "A more humane way of slaughtering chickens might get EU approval". New Scientist. Archived from the original on 24 June 2018. Retrieved 24 June 2018.
  61. ^ "What is CO2 stunning?". RSPCA. Archived from the original on 9 April 2014.
  62. ^ Campbell A (10 March 2018). "Humane execution and the fear of the tumbril". New Scientist. Archived from the original on 24 June 2018. Retrieved 24 June 2018.
  63. ^ Nordestgaard BG, Rostgaard J (February 1985). "Critical-point drying versus freeze drying for scanning electron microscopy: a quantitative and qualitative study on isolated hepatocytes". Journal of Microscopy. 137 (Pt 2): 189–207. doi:10.1111/j.1365-2818.1985.tb02577.x. PMID 3989858. S2CID 32065173.
  64. ^ "Types of Fire Extinguishers". The Fire Safety Advice Centre. Archived from the original on 28 June 2021. Retrieved 28 June 2021.
  65. ^ National Fire Protection Association Code 12.
  66. ^ Carbon Dioxide as a Fire Suppressant: Examining the Risks, US EPA. 2000.
  67. ^ "Appendix A: CO2 for use in enhanced oil recovery (EOR)". Accelerating the uptake of CCS: industrial use of captured carbon dioxide. Global CCS Institute. 20 December 2011. Archived from the original on 28 April 2017. Retrieved 2 January 2017.
  68. ^ Austell JM (2005). "CO2 for Enhanced Oil Recovery Needs – Enhanced Fiscal Incentives". Exploration & Production: The Oil & Gas Review. Archived from the original on 7 February 2012. Retrieved 28 September 2007.
  69. ^ "Enhanced coal bed methane recovery". ETH Zurich. 31 August 2006. Archived from the original on 6 July 2011.
  70. ^ Clayton M (11 January 2006). "Algae – like a breath mint for smokestacks". The Christian Science Monitor. Archived from the original on 14 September 2008. Retrieved 11 October 2007.
  71. ^ Atsumi S, Higashide W, Liao JC (December 2009). "Direct photosynthetic recycling of carbon dioxide to isobutyraldehyde". Nature Biotechnology. 27 (12): 1177–1180. doi:10.1038/nbt.1586. PMID 19915552. S2CID 1492698.
  72. ^ Cobb S, Badiani V, Dharani A, Wagner A, Zacarias S, Oliveira AR, et al. (28 February 2022). "Fast CO2 hydration kinetics impair heterogeneous but improve enzymatic CO2 reduction catalysis". Nature Chemistry. 14 (4): 417–424. Bibcode:2022NatCh..14..417C. doi:10.1038/s41557-021-00880-2. ISSN 1755-4349. PMC 7612589. PMID 35228690. S2CID 247160910.
  73. ^ Edwardes Moore E, Cobb SJ, Coito AM, Oliveira AR, Pereira IA, Reisner E (January 2022). "Understanding the local chemical environment of bioelectrocatalysis". Proceedings of the National Academy of Sciences of the United States of America. 119 (4): e2114097119. Bibcode:2022PNAS..11914097E. doi:10.1073/pnas.2114097119. PMC 8795565. PMID 35058361.
  74. ^ "Clean Way To Turn CO2 Into Fuel Inspired by Nature". Applied Sciences from Technology Networks. 1 March 2022. Retrieved 2 March 2022.
  75. ^ "The Coca-Cola Company Announces Adoption of HFC-Free Insulation in Refrigeration Units to Combat Global Warming". The Coca-Cola Company. 5 June 2006. Archived from the original on 1 November 2013. Retrieved 11 October 2007.
  76. ^ "Modine reinforces its CO2 research efforts". R744.com. 28 June 2007. Archived from the original on 10 February 2008.
  77. ^ TCE, the Chemical Engineer. Institution of Chemical Engineers. 1990. Archived from the original on 17 August 2021. Retrieved 2 June 2020.
  78. ^ a b "AVMA guidelines for the euthanasia of animals: 2020 Edition" (PDF). American Veterinary Medical Association. 2020. Archived (PDF) from the original on 1 February 2014. Retrieved 13 August 2021.
  79. ^ "Monthly Average Mauna Loa CO2". National Oceanic and Atmospheric Administration Earth System Research Laboratory, Global Monitoring Division. Archived from the original on 16 March 2007. Retrieved 19 December 2020.
  80. ^ National Oceanic & Atmospheric Administration (NOAA) – Earth System Research Laboratory (ESRL), Trends in Carbon Dioxide: Globally averaged marine surface monthly mean data. Archived 1 October 2021 at the Wayback Machine Values given are dry air mole fractions expressed in parts per million (ppm). For an ideal gas mixture this is equivalent to parts per million by volume (ppmv).
  81. ^ Pashley A (10 March 2016). "CO2 levels make largest recorded annual leap, Noaa data shows". The Guardian. Archived from the original on 14 March 2016. Retrieved 14 March 2016.
  82. ^ George K, Ziska LH, Bunce JA, Quebedeaux B (2007). "Elevated atmospheric CO2 concentration and temperature across an urban–rural transect". Atmospheric Environment. 41 (35): 7654–7665. Bibcode:2007AtmEn..41.7654G. doi:10.1016/j.atmosenv.2007.08.018. Archived from the original on 15 October 2019. Retrieved 12 September 2019.
  83. ^ Pisoft, Petr (25 May 2021). "Stratospheric contraction caused by increasing greenhouse gases". Environmental Research Letters. 16 (6): 064038. Bibcode:2021ERL....16f4038P. doi:10.1088/1748-9326/abfe2b.
  84. ^ Li AH (2016). "Hopes of Limiting Global Warming?". China Perspectives. 2016: 49–54. doi:10.4000/chinaperspectives.6924. Archived from the original on 12 May 2021. Retrieved 5 May 2021.
  85. ^ a b "The NOAA Annual Greenhouse Gas Index (AGGI) – An Introduction". NOAA Global Monitoring Laboratory/Earth System Research Laboratories. Archived from the original on 27 November 2020. Retrieved 18 December 2020.
  86. ^ a b Friedlingstein P, Jones MW, O'sullivan M, Andrew RM, Hauck J, Peters GP, et al. (2019). "Global Carbon Budget 2019". Earth System Science Data. 11 (4): 1783–1838. Bibcode:2019ESSD...11.1783F. doi:10.5194/essd-11-1783-2019..
  87. ^ "Global Warming Frequently Asked Questions". Climate.gov. NOAA. Archived from the original on 11 January 2017.
  88. ^ Gerlach TM (4 June 1991). "Present-day CO2 emissions from volcanoes". Eos, Transactions, American Geophysical Union. 72 (23): 249, 254–255. Bibcode:1991EOSTr..72..249.. doi:10.1029/90EO10192.
  89. ^ Buis A, Ramsayer K, Rasmussen C (12 November 2015). "A breathing planet, off balance". NASA. Archived from the original on 14 November 2015. Retrieved 13 November 2015.
  90. ^ "Audio (66:01) – NASA News Conference – Carbon & Climate Telecon". NASA. 12 November 2015. Archived from the original on 1 April 2019. Retrieved 12 November 2015.
  91. ^ St Fleur N (10 November 2015). "Atmospheric Greenhouse Gas Levels Hit Record, Report Says". The New York Times. Archived from the original on 11 November 2015. Retrieved 11 November 2015.
  92. ^ Ritter K (9 November 2015). "UK: In 1st, global temps average could be 1 degree C higher". Associated Press. Archived from the original on 17 November 2015. Retrieved 11 November 2015.
  93. ^ Atkins P, de Paula J (2006). Atkins' Physical Chemistry (8th ed.). W. H. Freeman. p. 462. ISBN 978-0-7167-8759-4.
  94. ^ "Carbon Dioxide Absorbs and Re-emits Infrared Radiation". UCAR Center for Science Education. 2012. Archived from the original on 21 September 2017. Retrieved 9 September 2017.
  95. ^ Archer D (15 March 2005). "How long will global warming last?". RealClimate. Archived from the original on 4 March 2021. Retrieved 5 March 2021.
  96. ^ Archer D (2009). "Atmospheric lifetime of fossil fuel carbon dioxide". Annual Review of Earth and Planetary Sciences. 37 (1): 117–34. Bibcode:2009AREPS..37..117A. doi:10.1146/annurev.earth.031208.100206. hdl:2268/12933. Archived from the original on 24 February 2021. Retrieved 7 March 2021.
  97. ^ Joos F, Roth R, Fuglestvedt JS, Peters GP, Enting IG, Von Bloh W, et al. (2013). "Carbon dioxide and climate impulse response functions for the computation of greenhouse gas metrics: A multi-model analysis". Atmospheric Chemistry and Physics. 13 (5): 2793–2825. doi:10.5194/acpd-12-19799-2012. Archived from the original on 22 July 2020. Retrieved 7 March 2021.
  98. ^ "Figure 8.SM.4" (PDF). Intergovernmental Panel on Climate Change Fifth Assessment Report. p. 8SM-16. Archived (PDF) from the original on 24 March 2021. Retrieved 7 March 2021.
  99. ^ Genthon G, Barnola JM, Raynaud D, Lorius C, Jouzel J, Barkov NI, et al. (1987). "Vostok ice core: climatic response to CO2 and orbital forcing changes over the last climatic cycle". Nature. 329 (6138): 414–418. Bibcode:1987Natur.329..414G. doi:10.1038/329414a0. S2CID 4333499.
  100. ^ "Climate and CO2 in the Atmosphere". Archived from the original on 6 October 2018. Retrieved 10 October 2007.
  101. ^ Berner RA, Kothavala Z (2001). "GEOCARB III: A revised model of atmospheric CO2 over Phanerozoic Time" (PDF). American Journal of Science. 301 (2): 182–204. Bibcode:2001AmJS..301..182B. CiteSeerX 10.1.1.393.582. doi:10.2475/ajs.301.2.182. Archived (PDF) from the original on 4 September 2011. Retrieved 15 February 2008.
  102. ^ van Gardingen PR, Grace J, Jeffree CE, Byari SH, Miglietta F, Raschi A, Bettarini I (1997). "Long-term effects of enhanced CO2 concentrations on leaf gas exchange: research opportunities using CO2 springs". In Raschi A, Miglietta F, Tognetti R, van Gardingen PR (eds.). Plant responses to elevated CO2: Evidence from natural springs. Cambridge: Cambridge University Press. pp. 69–86. ISBN 978-0-521-58203-2.
  103. ^ Martini M (1997). "CO2 emissions in volcanic areas: case histories and hazards". In Raschi A, Miglietta F, Tognetti R, van Gardingen PR (eds.). Plant responses to elevated CO2: Evidence from natural springs. Cambridge: Cambridge University Press. pp. 69–86. ISBN 978-0-521-58203-2.
  104. ^ Doney SC, Levine NM (29 November 2006). "How Long Can the Ocean Slow Global Warming?". Oceanus. Archived from the original on 4 January 2008. Retrieved 21 November 2007.
  105. ^ Nienhuis S, Palmer AR, Harley CD (August 2010). "Elevated CO2 affects shell dissolution rate but not calcification rate in a marine snail". Proceedings. Biological Sciences. 277 (1693): 2553–2558. doi:10.1098/rspb.2010.0206. PMC 2894921. PMID 20392726.
  106. ^ Gattuso JP, Frankignoulle M, Bourge I, Romaine S, Buddemeier RW (1998). "Effect of calcium carbonate saturation of seawater on coral calcification". Global and Planetary Change. 18 (1–2): 37–46. Bibcode:1998GPC....18...37G. doi:10.1016/S0921-8181(98)00035-6.
  107. ^ Gattuso JP, Allemand D, Frankignoulle M (1999). "Photosynthesis and calcification at cellular, organismal and community levels in coral reefs: a review on interactions and control by carbonate chemistry". American Zoologist. 39: 160–183. doi:10.1093/icb/39.1.160.
  108. ^ Langdon C, Atkinson MJ (2005). "Effect of elevated pCO2 on photosynthesis and calcification of corals and interactions with seasonal change in temperature/irradiance and nutrient enrichment". Journal of Geophysical Research. 110 (C09S07): C09S07. Bibcode:2005JGRC..110.9S07L. doi:10.1029/2004JC002576.
  109. ^ Riebesell U, Zondervan I, Rost B, Tortell PD, Zeebe RE, Morel FM (September 2000). "Reduced calcification of marine plankton in response to increased atmospheric CO2" (PDF). Nature. 407 (6802): 364–367. Bibcode:2000Natur.407..364R. doi:10.1038/35030078. PMID 11014189. S2CID 4426501. Archived (PDF) from the original on 25 August 2020. Retrieved 24 August 2020.
  110. ^ Zondervan I, Zeebe RE, Rost B, Rieblesell U (2001). "Decreasing marine biogenic calcification: a negative feedback on rising atmospheric CO2" (PDF). Global Biogeochemical Cycles. 15 (2): 507–516. Bibcode:2001GBioC..15..507Z. doi:10.1029/2000GB001321. Archived (PDF) from the original on 21 July 2018. Retrieved 15 October 2019.
  111. ^ Zondervan I, Rost B, Rieblesell U (2002). "Effect of CO2 concentration on the PIC/POC ratio in the coccolithophore Emiliania huxleyi grown under light limiting conditions and different day lengths" (PDF). Journal of Experimental Marine Biology and Ecology. 272 (1): 55–70. doi:10.1016/S0022-0981(02)00037-0. Archived (PDF) from the original on 19 July 2018. Retrieved 7 November 2018.
  112. ^ Delille B, Harlay J, Zondervan I, Jacquet S, Chou L, Wollast R, et al. (2005). "Response of primary production and calcification to changes of pCO2 during experimental blooms of the coccolithophorid Emiliania huxleyi". Global Biogeochemical Cycles. 19 (2): GB2023. Bibcode:2005GBioC..19.2023D. doi:10.1029/2004GB002318.
  113. ^ Kuffner IB, Andersson AJ, Jokiel PL, Rodgers KU, Mackenzie FT (2007). "Decreased abundance of crustose coralline algae due to ocean acidification". Nature Geoscience. 1 (2): 114–117. Bibcode:2008NatGe...1..114K. doi:10.1038/ngeo100.
  114. ^ Phillips G, Branagan C (13 September 2007). "Ocean Acidification – The BIG global warming story". ABC TV Science: Catalyst. Australian Broadcasting Corporation. Archived from the original on 11 October 2007. Retrieved 18 September 2007.
  115. ^ Gazeau F, Quiblier C, Jansen JM, Gattuso JP, Middelburg JJ, Heip CH (2007). "Impact of elevated CO
    2
    on shellfish calcification". Geophysical Research Letters. 34 (7): L07603. Bibcode:2007GeoRL..34.7603G. CiteSeerX 10.1.1.326.1630. doi:10.1029/2006GL028554. hdl:20.500.11755/a8941c6a-6d0b-43d5-ba0d-157a7aa05668. S2CID 130190489.
  116. ^ Comeau S, Gorsky G, Jeffree R, Teyssié JL, Gattuso JP (2009). "Impact of ocean acidification on a key Arctic pelagic mollusc (Limacina helicina)". Biogeosciences. 6 (9): 1877–1882. Bibcode:2009BGeo....6.1877C. doi:10.5194/bg-6-1877-2009.
  117. ^ Duana Z, Sun R (2003). "An improved model calculating CO2 solubility in pure water and aqueous NaCl solutions from 273 to 533 K and from 0 to 2000 bar". Chemical Geology. 193 (3–4): 257–271. Bibcode:2003ChGeo.193..257D. doi:10.1016/S0009-2541(02)00263-2.
  118. ^ Cai WJ, Chen L, Chen B, Gao Z, Lee SH, Chen J, et al. (July 2010). "Decrease in the CO2 uptake capacity in an ice-free Arctic Ocean basin". Science. 329 (5991): 556–559. Bibcode:2010Sci...329..556C. doi:10.1126/science.1189338. PMID 20651119. S2CID 206526452.
  119. ^ Garrison T (2004). Oceanography: An Invitation to Marine Science. Thomson Brooks. p. 125. ISBN 978-0-534-40887-9.
  120. ^ Ries JB, Cohen AL, McCorkle DC (2009). "Marine calcifiers exhibit mixed responses to CO2-induced ocean acidification". Geology. 37 (12): 1131–1134. Bibcode:2009Geo....37.1131R. doi:10.1130/G30210A.1.
  121. ^ "State of the Science FACT SHEET: Ocean Acidification" (PDF). National Oceanic and Atmospheric Administration. May 2008. Archived (PDF) from the original on 10 October 2022. Retrieved 2 October 2021.
  122. ^ Pachauri RK, Reisinger A (eds.). "Climate Change 2007: Synthesis Report". Intergovernmental Panel on Climate Change (IPCC). Geneva, Switzerland. Archived from the original on 2 November 2018.
  123. ^ "PMEL Ocean Acidification Home Page". National Oceanic and Atmospheric Administration. Archived from the original on 9 December 2013. Retrieved 14 January 2014.
  124. ^ Lupton J, Lilley M, Butterfield D, Evans L, Embley R, Olson E, et al. (2004). "Liquid Carbon Dioxide Venting at the Champagne Hydrothermal Site, NW Eifuku Volcano, Mariana Arc". American Geophysical Union. 2004 (Fall Meeting). V43F–08. Bibcode:2004AGUFM.V43F..08L.
  125. ^ Inagaki F, Kuypers MM, Tsunogai U, Ishibashi J, Nakamura K, Treude T, et al. (September 2006). "Microbial community in a sediment-hosted CO2 lake of the southern Okinawa Trough hydrothermal system". Proceedings of the National Academy of Sciences of the United States of America. 103 (38): 14164–14169. Bibcode:2006PNAS..10314164I. doi:10.1073/pnas.0606083103. PMC 1599929. PMID 16959888. Videos can be downloaded at "Supporting Information". Archived from the original on 19 October 2018.
  126. ^ Dhingra A, Portis AR, Daniell H (April 2004). "Enhanced translation of a chloroplast-expressed RbcS gene restores small subunit levels and photosynthesis in nuclear RbcS antisense plants". Proceedings of the National Academy of Sciences of the United States of America. 101 (16): 6315–6320. Bibcode:2004PNAS..101.6315D. doi:10.1073/pnas.0400981101. PMC 395966. PMID 15067115. (Rubisco) is the most prevalent enzyme on this planet, accounting for 30–50% of total soluble protein in the chloroplast
  127. ^ Falkowski P, Knoll AH (1 January 2007). Evolution of primary producers in the sea. Elsevier, Academic Press. ISBN 978-0-12-370518-1. OCLC 845654016.
  128. ^ Blom TJ, Straver WA, Ingratta FJ, Khosla S, Brown W (December 2002). "Carbon Dioxide In Greenhouses". Archived from the original on 29 April 2019. Retrieved 12 June 2007.
  129. ^ Ainsworth EA (2008). "Rice production in a changing climate: a meta-analysis of responses to elevated carbon dioxide and elevated ozone concentration" (PDF). Global Change Biology. 14 (7): 1642–1650. Bibcode:2008GCBio..14.1642A. doi:10.1111/j.1365-2486.2008.01594.x. S2CID 19200429. Archived from the original (PDF) on 19 July 2011.
  130. ^ Long SP, Ainsworth EA, Leakey AD, Nösberger J, Ort DR (June 2006). "Food for thought: lower-than-expected crop yield stimulation with rising CO2 concentrations" (PDF). Science. 312 (5782): 1918–1921. Bibcode:2006Sci...312.1918L. CiteSeerX 10.1.1.542.5784. doi:10.1126/science.1114722. PMID 16809532. S2CID 2232629. Archived (PDF) from the original on 20 October 2016. Retrieved 27 October 2017.
  131. ^ Woodward F, Kelly C (1995). "The influence of CO2 concentration on stomatal density". New Phytologist. 131 (3): 311–327. doi:10.1111/j.1469-8137.1995.tb03067.x.
  132. ^ Drake BG, Gonzalez-Meler MA, Long SP (June 1997). "MORE EFFICIENT PLANTS: A Consequence of Rising Atmospheric CO2?". Annual Review of Plant Physiology and Plant Molecular Biology. 48 (1): 609–639. doi:10.1146/annurev.arplant.48.1.609. PMID 15012276. S2CID 33415877.
  133. ^ Loladze I (2002). "Rising atmospheric CO2 and human nutrition: toward globally imbalanced plant stoichiometry?". Trends in Ecology & Evolution. 17 (10): 457–461. doi:10.1016/S0169-5347(02)02587-9. S2CID 16074723.
  134. ^ Coviella CE, Trumble JT (1999). "Effects of Elevated Atmospheric Carbon Dioxide on Insect-Plant Interactions". Conservation Biology. 13 (4): 700–712. doi:10.1046/j.1523-1739.1999.98267.x. JSTOR 2641685. S2CID 52262618.
  135. ^ Davey MP, Harmens H, Ashenden TW, Edwards R, Baxter R (2007). "Species-specific effects of elevated CO2 on resource allocation in Plantago maritima and Armeria maritima". Biochemical Systematics and Ecology. 35 (3): 121–129. doi:10.1016/j.bse.2006.09.004.
  136. ^ Davey MP, Bryant DN, Cummins I, Ashenden TW, Gates P, Baxter R, Edwards R (August 2004). "Effects of elevated CO2 on the vasculature and phenolic secondary metabolism of Plantago maritima". Phytochemistry. 65 (15): 2197–2204. doi:10.1016/j.phytochem.2004.06.016. PMID 15587703.
  137. ^ "Global Environment Division Greenhouse Gas Assessment Handbook – A Practical Guidance Document for the Assessment of Project-level Greenhouse Gas Emissions". World Bank. Archived from the original on 3 June 2016. Retrieved 10 November 2007.
  138. ^ Luyssaert S, Schulze ED, Börner A, Knohl A, Hessenmöller D, Law BE, et al. (September 2008). "Old-growth forests as global carbon sinks" (PDF). Nature. 455 (7210): 213–215. Bibcode:2008Natur.455..213L. doi:10.1038/nature07276. PMID 18784722. S2CID 4424430.
  139. ^ Falkowski P, Scholes RJ, Boyle E, Canadell J, Canfield D, Elser J, et al. (October 2000). "The global carbon cycle: a test of our knowledge of earth as a system". Science. 290 (5490): 291–296. Bibcode:2000Sci...290..291F. doi:10.1126/science.290.5490.291. PMID 11030643. S2CID 1779934.
  140. ^ a b Friedman D. "Toxicity of Carbon Dioxide Gas Exposure, CO2 Poisoning Symptoms, Carbon Dioxide Exposure Limits, and Links to Toxic Gas Testing Procedures". InspectAPedia. Archived from the original on 28 September 2009.
  141. ^ "CarbonTracker CT2011_oi (Graphical map of CO2)". esrl.noaa.gov. Archived from the original on 13 February 2021. Retrieved 20 April 2007.
  142. ^ "Carbon Dioxide as a Fire Suppressant: Examining the Risks". U.S. Environmental Protection Agency. Archived from the original on 2 October 2015.
  143. ^ "Volcano Under the City". A NOVA Production by Bonne Pioche and Greenspace for WGBH/Boston. Public Broadcasting System. 1 November 2005. Archived from the original on 5 April 2011..
  144. ^ Glatte Jr HA, Motsay GJ, Welch BE (1967). Carbon Dioxide Tolerance Studies (Report). Brooks AFB, TX School of Aerospace Medicine Technical Report. SAM-TR-67-77. Archived from the original on 9 May 2008. Retrieved 2 May 2008.{{cite report}}: CS1 maint: unfit URL (link)
  145. ^ Lambertsen CJ (1971). Carbon Dioxide Tolerance and Toxicity (Report). IFEM Report. Environmental Biomedical Stress Data Center, Institute for Environmental Medicine, University of Pennsylvania Medical Center. No. 2-71. Archived from the original on 24 July 2011. Retrieved 2 May 2008.{{cite report}}: CS1 maint: unfit URL (link)
  146. ^ a b Satish U, Mendell MJ, Shekhar K, Hotchi T, Sullivan D, Streufert S, Fisk WJ (December 2012). "Is CO2 an indoor pollutant? Direct effects of low-to-moderate CO2 concentrations on human decision-making performance" (PDF). Environmental Health Perspectives. 120 (12): 1671–1677. doi:10.1289/ehp.1104789. PMC 3548274. PMID 23008272. Archived from the original (PDF) on 5 March 2016. Retrieved 11 December 2014.
  147. ^ a b Allen JG, MacNaughton P, Satish U, Santanam S, Vallarino J, Spengler JD (June 2016). "Associations of Cognitive Function Scores with Carbon Dioxide, Ventilation, and Volatile Organic Compound Exposures in Office Workers: A Controlled Exposure Study of Green and Conventional Office Environments". Environmental Health Perspectives. 124 (6): 805–812. doi:10.1289/ehp.1510037. PMC 4892924. PMID 26502459.
  148. ^ "Exposure Limits for Carbon Dioxide Gas – CO2 Limits". InspectAPedia.com. Archived from the original on 16 September 2018. Retrieved 19 October 2014.
  149. ^ Law J, Watkins S, Alexander D (2010). In-Flight Carbon Dioxide Exposures and Related Symptoms: Associations, Susceptibility and Operational Implications (PDF) (Report). NASA Technical Report. TP–2010–216126. Archived from the original (PDF) on 27 June 2011. Retrieved 26 August 2014.
  150. ^ Schaefer KE, Douglas WH, Messier AA, Shea ML, Gohman PA (1979). "Effect of prolonged exposure to 0.5% CO2 on kidney calcification and ultrastructure of lungs". Undersea Biomedical Research. 6 (Suppl): S155–S161. PMID 505623. Archived from the original on 19 October 2014. Retrieved 19 October 2014.
  151. ^ Du B, Tandoc MC, Mack ML, Siegel JA (November 2020). "Indoor CO2 concentrations and cognitive function: A critical review". Indoor Air. 30 (6): 1067–1082. doi:10.1111/ina.12706. PMID 32557862. S2CID 219915861.
  152. ^ Kaplan L (4 June 2019). "Ask the doc: Does my helmet make me stupid? - RevZilla". www.revzilla.com. Archived from the original on 22 May 2021. Retrieved 22 May 2021.
  153. ^ Brühwiler PA, Stämpfli R, Huber R, Camenzind M (September 2005). "CO2 and O2 concentrations in integral motorcycle helmets". Applied Ergonomics. 36 (5): 625–633. doi:10.1016/j.apergo.2005.01.018. PMID 15893291.
  154. ^ Allen JG, MacNaughton P, Satish U, Santanam S, Vallarino J, Spengler JD (June 2016). "Associations of Cognitive Function Scores with Carbon Dioxide, Ventilation, and Volatile Organic Compound Exposures in Office Workers: A Controlled Exposure Study of Green and Conventional Office Environments". Environmental Health Perspectives. 124 (6): 805–812. doi:10.1289/ehp.1510037. PMC 4892924. PMID 26502459.
  155. ^ Romm J (26 October 2015). "Exclusive: Elevated CO2 Levels Directly Affect Human Cognition, New Harvard Study Shows". ThinkProgress. Archived from the original on 9 October 2019. Retrieved 14 October 2019.
  156. ^ "Three die in dry-ice incident at Moscow pool party". BBC News. 29 February 2020. Archived from the original on 29 February 2020. The victims were connected to Instagram influencer Yekaterina Didenko.
  157. ^ Rettner R (2 August 2018). "A Woman Died from Dry Ice Fumes. Here's How It Can Happen". livescience.com. Archived from the original on 22 May 2021. Retrieved 22 May 2021.
  158. ^ a b "ABG (Arterial Blood Gas)". Brookside Associates. Archived from the original on 12 August 2017. Retrieved 2 January 2017.
  159. ^ "How much carbon dioxide do humans contribute through breathing?". EPA.gov. Archived from the original on 2 February 2011. Retrieved 30 April 2009.
  160. ^ Henrickson C (2005). Chemistry. Cliffs Notes. ISBN 978-0-7645-7419-1.
  161. ^ a b c d "Carbon dioxide". solarnavigator.net. Archived from the original on 14 September 2008. Retrieved 12 October 2007.
  162. ^ Battisti-Charbonney, A.; Fisher, J.; Duffin, J. (15 June 2011). "The cerebrovascular response to carbon dioxide in humans". J. Physiol. 589 (12): 3039–3048. doi:10.1113/jphysiol.2011.206052. PMC 3139085. PMID 21521758.
  163. ^ Patel, S.; Miao, J.H.; Yetiskul, E.; Anokhin, A.; Majmunder, S.H. (2022). "Physiology, Carbon Dioxide Retention". National Library of Medicine. National Center for Biotechnology Information, NIH. PMID 29494063. Retrieved 20 August 2022.
  164. ^ Wilmshurst, Peter (1998). "ABC of oxygen". BMJ. 317 (7164): 996–999. doi:10.1136/bmj.317.7164.996. PMC 1114047. PMID 9765173.

Further reading

  • Seppänen OA, Fisk WJ, Mendell MJ (December 1999). "Association of ventilation rates and CO2 concentrations with health and other responses in commercial and institutional buildings" (PDF). Indoor Air. 9 (4): 226–252. doi:10.1111/j.1600-0668.1999.00003.x. PMID 10649857. Archived from the original (PDF) on 27 December 2016.
  • Shendell DG, Prill R, Fisk WJ, Apte MG, Blake D, Faulkner D (October 2004). "Associations between classroom CO2 concentrations and student attendance in Washington and Idaho" (PDF). Indoor Air. 14 (5): 333–341. doi:10.1111/j.1600-0668.2004.00251.x. hdl:2376/5954. PMID 15330793. Archived from the original (PDF) on 27 December 2016.
  • Soentgen J (February 2014). "Hot air: The science and politics of CO2". Global Environment. 7 (1): 134–171. doi:10.3197/197337314X13927191904925. Archived from the original on 1 October 2021. Retrieved 5 May 2021.
  • Good plant design and operation for onshore carbon capture installations and onshore pipelines: a recommended practice guidance document. Global CCS Institute. Energy Institute and Global Carbon Capture and Storage Institute. 1 September 2010. Archived from the original on 7 November 2018. Retrieved 2 January 2018. This new title is an essential guide for engineers, managers, procurement specialists and designers working on global carbon capture and storage projects.

  • International Chemical Safety Card 0021
  • Current global map of carbon dioxide concentration
  • CDC – NIOSH Pocket Guide to Chemical Hazards – Carbon Dioxide
  • CO2 Carbon Dioxide Properties, Uses, Applications
  • Dry Ice information
  • Trends in Atmospheric Carbon Dioxide (NOAA)
  • "A War Gas That Saves Lives". Popular Science, June 1942, pp. 53–57.
  • Reactions, Thermochemistry, Uses, and Function of Carbon Dioxide
  • Carbon Dioxide – Part One and Carbon Dioxide – Part Two at The Periodic Table of Videos (University of Nottingham)