Hydrogen sulfide

Hydrogen sulfide is the chemical compound with the formula H
2
S
. It is a colorless chalcogen hydride gas with the characteristic foul odor of rotten eggs. It is very poisonous, corrosive, and flammable.[9]

Hydrogen sulfide
Ball-and-stick model of hydrogen sulfide
Spacefill model of hydrogen sulfide
Names
Systematic IUPAC name
Hydrogen sulfide[1]
Other names
  • Dihydrogen monosulfide
  • Sour Gas
  • Dihydrogen sulfide
  • Sewer gas
  • Sulfane
  • Sulfurated hydrogen
  • Sulfureted hydrogen
  • Sulfuretted hydrogen
  • Sulfur hydride
  • Hydrosulfuric acid
  • Hydrothionic acid
  • Thiohydroxic acid
  • Sulfhydric acid
Identifiers
3D model (JSmol)
3DMet
Beilstein Reference
3535004
ChEBI
ChEMBL
ChemSpider
ECHA InfoCard 100.029.070
EC Number
  • 231-977-3
Gmelin Reference
303
KEGG
MeSH Hydrogen+sulfide
RTECS number
  • MX1225000
UNII
UN number 1053
CompTox Dashboard (EPA)
Properties
H2S
Molar mass 34.08 g·mol−1
Appearance Colorless gas
Odor Pungent, like that of rotten eggs
Density 1.363 g dm−3
Melting point −82 °C (−116 °F; 191 K)
Boiling point −60 °C (−76 °F; 213 K)
4 g dm−3 (at 20 °C)
Vapor pressure 1740 kPa (at 21 °C)
Acidity (pKa) 7.0[2][3]
Conjugate acid Sulfonium
Conjugate base Bisulfide
25.5·10−6 cm3/mol
1.000644 (0 °C)[4]
Structure
C2v
Molecular shape
Bent
0.97 D
Thermochemistry
1.003 J K−1 g−1
Std molar
entropy (So298)
206 J mol−1 K−1[5]
−21 kJ mol−1[5]
Hazards
Main hazards Flammable and highly toxic
EU classification (DSD) (outdated)
F+ T+ N
R-phrases (outdated) R12, R26, R50
S-phrases (outdated) (S1/2), S9, S16, S36, S38, S45, S61
NFPA 704 (fire diamond)
4
4
0
Flash point −82.4 °C (−116.3 °F; 190.8 K) [6]
Autoignition
temperature
232 °C (450 °F; 505 K)
Explosive limits 4.3–46%
Lethal dose or concentration (LD, LC):
  • 713 ppm (rat, 1 hr)
  • 673 ppm (mouse, 1 hr)
  • 634 ppm (mouse, 1 hr)
  • 444 ppm (rat, 4 hr)[7]
  • 600 ppm (human, 30 min)
  • 800 ppm (human, 5 min)[7]
NIOSH (US health exposure limits):
PEL (Permissible)
C 20 ppm; 50 ppm [10-minute maximum peak][8]
REL (Recommended)
C 10 ppm (15 mg/m3) [10-minute][8]
IDLH (Immediate danger)
100 ppm[8]
Related compounds
Related hydrogen chalcogenides
  • Water
  • Hydrogen selenide
  • Hydrogen telluride
  • Hydrogen polonide
  • Hydrogen disulfide
  • Sulfanyl
Related compounds
Phosphine
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
N verify (what is YN ?)
Infobox references

Hydrogen sulfide is often produced from the microbial breakdown of organic matter in the absence of oxygen gas, such as in swamps and sewers; this process is commonly known as anaerobic digestion which is done by sulfate-reducing microorganisms. H
2
S
also occurs in volcanic gases, natural gas, and in some sources of well water.[10] The human body produces small amounts of H
2
S
and uses it as a signaling molecule.[11]

Swedish chemist Carl Wilhelm Scheele is credited with having discovered the chemical composition of hydrogen sulfide in 1777.

The British English spelling of this compound is hydrogen sulphide, but this spelling is not recommended by the International Union of Pure and Applied Chemistry (IUPAC) or the Royal Society of Chemistry.

Properties

Hydrogen sulfide is slightly denser than air; a mixture of H
2
S
and air can be explosive. Hydrogen sulfide burns in oxygen with a blue flame to form sulfur dioxide (SO
2
) and water. In general, hydrogen sulfide acts as a reducing agent, especially in the presence of base, which forms SH.

At high temperatures or in the presence of catalysts, sulfur dioxide reacts with hydrogen sulfide to form elemental sulfur and water. This reaction is exploited in the Claus process, an important industrial method to dispose of hydrogen sulfide.

Hydrogen sulfide is slightly soluble in water and acts as a weak acid (pKa = 6.9 in 0.01–0.1 mol/litre solutions at 18 °C), giving the hydrosulfide ion HS
(also written SH
). Hydrogen sulfide and its solutions are colorless. When exposed to air, it slowly oxidizes to form elemental sulfur, which is not soluble in water. The sulfide anion S2−
is not formed in aqueous solution.[12]

Hydrogen sulfide reacts with metal ions to form metal sulfides, which are insoluble, often dark colored solids. Lead(II) acetate paper is used to detect hydrogen sulfide because it readily converts to lead(II) sulfide, which is black. Treating metal sulfides with strong acid often liberates hydrogen sulfide.

At pressures above 90 GPa (gigapascal), hydrogen sulfide becomes a metallic conductor of electricity. When cooled below a critical temperature this high-pressure phase exhibits superconductivity. The critical temperature increases with pressure, ranging from 23 K at 100 GPa to 150 K at 200 GPa.[13] If hydrogen sulfide is pressurized at higher temperatures, then cooled, the critical temperature reaches 203 K (−70 °C), the highest accepted superconducting critical temperature as of 2015. By substituting a small part of sulfur with phosphorus and using even higher pressures, it has been predicted that it may be possible to raise the critical temperature to above 0 °C (273 K) and achieve room-temperature superconductivity.[14]

Production

Hydrogen sulfide is most commonly obtained by its separation from sour gas, which is natural gas with a high content of H
2
S
. It can also be produced by treating hydrogen with molten elemental sulfur at about 450 °C. Hydrocarbons can serve as a source of hydrogen in this process.[15]

Sulfate-reducing (resp. sulfur-reducing) bacteria generate usable energy under low-oxygen conditions by using sulfates (resp. elemental sulfur) to oxidize organic compounds or hydrogen; this produces hydrogen sulfide as a waste product.

A standard lab preparation is to treat ferrous sulfide with a strong acid in a Kipp generator:

FeS + 2 HCl → FeCl2 + H2S

For use in qualitative inorganic analysis, thioacetamide is used to generate H
2
S
:

CH3C(S)NH2 + H2O → CH3C(O)NH2 + H2S

Many metal and nonmetal sulfides, e.g. aluminium sulfide, phosphorus pentasulfide, silicon disulfide liberate hydrogen sulfide upon exposure to water:[16]

6 H2O + Al2S3 → 3 H2S + 2 Al(OH)3

This gas is also produced by heating sulfur with solid organic compounds and by reducing sulfurated organic compounds with hydrogen.

Water heaters can aid the conversion of sulfate in water to hydrogen sulfide gas. This is due to providing a warm environment sustainable for sulfur bacteria and maintaining the reaction which interacts between sulfate in the water and the water heater anode, which is usually made from magnesium metal.[17]

Biosynthesis in the body

Hydrogen sulfide can be generated in cells via enzymatic or non enzymatic pathway. H
2
S
in the body acts as a gaseous signaling molecule which is known to inhibit Complex IV of the mitochondrial electron transport chain which effectively reduces ATP generation and biochemical activity within cells.[18] Three enzymes are known to synthesize H
2
S
: cystathionine γ-lyase (CSE), cystathionine β-synthetase (CBS) and 3-mercaptopyruvate sulfurtransferase (3-MST).[19] These enzymes have been identified in a breadth of biological cells and tissues, and their activity has been observed to be induced by a number of disease states.[20] It is becoming increasingly clear that H
2
S
is an important mediator of a wide range of cell functions in health and in disease.[19] CBS and CSE are the main proponents of H
2
S
biogenesis, which follows the trans-sulfuration pathway.[21] These enzymes are characterized by the transfer of a sulfur atom from methionine to serine to form a cysteine molecule.[21] 3-MST also contributes to hydrogen sulfide production by way of the cysteine catabolic pathway.[20][21] Dietary amino acids, such as methionine and cysteine serve as the primary substrates for the transulfuration pathways and in the production of hydrogen sulfide. Hydrogen sulfide can also be synthesized by non-enzymatic pathway, which is derived from proteins such as ferredoxins and Rieske proteins.[20]

H
2
S
has been shown to be involved in physiological processes like vasodilatation in animals, increasing seed germination and stress responses in plants.[18] Hydrogen sulfide signaling is also innately intertwined with physiological processes that are known to be moderated by reactive oxygen species (ROS) and reactive nitrogen species (RNS).[18] H
2
S
has been shown to interact with NO resulting in several different cellular effects, as well as the formation of a new signal called nitrosothiol.[18] Hydrogen Sulfide is also known to increase the levels of glutathione which acts to reduce or disrupt ROS levels in cells.[18] The field of H2S biology advanced from the environmental toxicology area to investigate the roles of endogenously produced H2S in physiological conditions and in various pathophysiological states.[22] According to a current classification, pathophysiological states with H2S overproduction (e.g. cancer, Down Syndrome) and pathophysiological states with H2S deficit (e.g. vascular disease) can be identified.[23] Although the understanding of H2S biology has significantly advanced over the last decade,[24][25][26] many questions remain, for instance related to the quantification of endogenous H2S levels[20]

Uses

Production of sulfur, thioorganic compounds, and alkali metal sulfides

The main use of hydrogen sulfide is as a precursor to elemental sulfur. Several organosulfur compounds are produced using hydrogen sulfide. These include methanethiol, ethanethiol, and thioglycolic acid.[15]

Upon combining with alkali metal bases, hydrogen sulfide converts to alkali hydrosulfides such as sodium hydrosulfide and sodium sulfide:

H2S + NaOH → NaSH + H2O
NaSH + NaOH → Na2S + H2O

These compounds are used in the paper making industry. Specifically, salts of SH break bonds between lignin and cellulose components of pulp in the Kraft process.[15]

Reversibly sodium sulfide in the presence of acids turns into hydrosulfides and hydrogen sulfide; this supplies hydrosulfides in organic solutions and is utilized in the production of thiophenol.[27]

Analytical chemistry

For well over a century hydrogen sulfide was important in analytical chemistry in the qualitative inorganic analysis of metal ions. In these analyses, heavy metal (and nonmetal) ions (e.g., Pb(II), Cu(II), Hg(II), As(III)) are precipitated from solution upon exposure to H
2
S
). The components of the resulting precipitate redissolve with some selectivity, and are thus identified.

Precursor to metal sulfides

As indicated above, many metal ions react with hydrogen sulfide to give the corresponding metal sulfides. This conversion is widely exploited. For example, gases or waters contaminated by hydrogen sulfide can be cleaned with metals, by forming metal sulfides. In the purification of metal ores by flotation, mineral powders are often treated with hydrogen sulfide to enhance the separation. Metal parts are sometimes passivated with hydrogen sulfide. Catalysts used in hydrodesulfurization are routinely activated with hydrogen sulfide, and the behavior of metallic catalysts used in other parts of a refinery is also modified using hydrogen sulfide.

Miscellaneous applications

Hydrogen sulfide is used to separate deuterium oxide, or heavy water, from normal water via the Girdler sulfide process.

Scientists from the University of Exeter discovered that cell exposure to small amounts of hydrogen sulfide gas can prevent mitochondrial damage. When the cell is stressed with disease, enzymes are drawn into the cell to produce small amounts of hydrogen sulfide. This study could have further implications on preventing strokes, heart disease and arthritis.[28]

A suspended animation-like state has been induced in rodents with the use of hydrogen sulfide, resulting in hypothermia with a concomitant reduction in metabolic rate. Also oxygen demand was reduced, thereby protecting against hypoxia. In addition, hydrogen sulfide has been shown to reduce inflammation in various situations.[29]

Occurrence

Deposit of sulfur on a rock, caused by volcanic gas

Small amounts of hydrogen sulfide occur in crude petroleum, but natural gas can contain up to 30%.[30] Volcanoes and some hot springs (as well as cold springs) emit some H
2
S
, where it probably arises via the hydrolysis of sulfide minerals, i.e. MS + H
2
O
→ MO + H
2
S
. Hydrogen sulfide can be present naturally in well water, often as a result of the action of sulfate-reducing bacteria. Hydrogen sulfide is created by the human body in small doses through bacterial breakdown of proteins containing sulfur in the intestinal tract. It is also produced in the mouth (halitosis).[31]

A portion of global H
2
S
emissions are due to human activity. By far the largest industrial source of H
2
S
is petroleum refineries: The hydrodesulfurization process liberates sulfur from petroleum by the action of hydrogen. The resulting H
2
S
is converted to elemental sulfur by partial combustion via the Claus process, which is a major source of elemental sulfur. Other anthropogenic sources of hydrogen sulfide include coke ovens, paper mills (using the Kraft process), tanneries and sewerage. H
2
S
arises from virtually anywhere where elemental sulfur comes in contact with organic material, especially at high temperatures. Depending on environmental conditions, it is responsible for deterioration of material through the action of some sulfur oxidizing microorganisms. It is called biogenic sulfide corrosion.

In 2011 it was reported that increased concentration of H
2
S
, possibly due to oil field practices, was observed in the Bakken formation crude and presented challenges such as "health and environmental risks, corrosion of wellbore, added expense with regard to materials handling and pipeline equipment, and additional refinement requirements".[32]

Besides living near a gas and oil drilling operations, ordinary citizens can be exposed to hydrogen sulfide by being near waste water treatment facilities, landfills and farms with manure storage. Exposure occurs through breathing contaminated air or drinking contaminated water.[33]

In municipal waste landfill sites, the burial of organic material rapidly leads to the production of anaerobic digestion within the waste mass and, with the humid atmosphere and relatively high temperature that accompanies biodegradation, biogas is produced as soon as the air within the waste mass has been reduced. If there is a source of sulfate bearing material, such as plasterboard or natural gypsum (calcium sulphate dihydrate), under anaerobic conditions sulfate reducing bacteria converts this to hydrogen sulfide. These bacteria cannot survive in air but the moist, warm, anaerobic conditions of buried waste that contains a high source of carbon in inert landfills, paper and glue used in the fabrication of products such as plasterboard can provide a rich source of carbon[34] is an excellent environment for the formation of hydrogen sulfide.

In industrial anaerobic digestion processes, such as waste water treatment or the digestion of organic waste from agriculture, hydrogen sulfide can be formed from the reduction of sulfate and the degradation of amino acids and proteins within organic compounds.[35] Sulfates are relatively non-inhibitory to methane forming bacteria but can be reduced to H2S by sulfate reducing bacteria, of which there are several genera.[36]

Removal from water

A number of processes has been designed to remove hydrogen sulfide from drinking water.[37]

Continuous chlorination
For levels up to 75 mg/L chlorine is used in the purification process as an oxidizing chemical to react with hydrogen sulfide. This reaction yields insoluble solid sulfur. Usually the chlorine used is in the form of sodium hypochlorite.[38]
Aeration
For concentrations of hydrogen sulfide less than 2 mg/L aeration is an ideal treatment process. Oxygen is added to water and a reaction between oxygen and hydrogen sulfide react to produce odorless sulfate.[39]
Nitrate addition
Calcium nitrate can be used to prevent hydrogen sulfide formation in wastewater streams.

Removal from fuel gases

Hydrogen sulfide is commonly found in raw natural gas and biogas. It is typically removed by amine gas treating technologies. In such processes, the hydrogen sulfide is first converted to an ammonium salt, whereas the natural gas is unaffected.

RNH2 + H2S RNH+
3
+ SH

The bisulfide anion is subsequently regenerated by heating of the amine sulfide solution. Hydrogen sulfide generated in this process is typically converted to elemental sulfur using the Claus Process.

Process flow diagram of a typical amine treating process used in petroleum refineries, natural gas processing plants and other industrial facilities.

Safety

Hydrogen sulfide is a highly toxic and flammable gas (flammable range: 4.3–46%). Being heavier than air, it tends to accumulate at the bottom of poorly ventilated spaces. Although very pungent at first, it quickly deadens the sense of smell (which smells like rotten egg(s)[40]), so victims may be unaware of its presence until it is too late. For safe handling procedures, a hydrogen sulfide safety data sheet (SDS) should be consulted.[41]

Toxicity

Hydrogen sulfide is a broad-spectrum poison, meaning that it can poison several different systems in the body, although the nervous system is most affected. The toxicity of H
2
S
is comparable with that of carbon monoxide.[42] It binds with iron in the mitochondrial cytochrome enzymes, thus preventing cellular respiration.

Since hydrogen sulfide occurs naturally in the body, the environment, and the gut, enzymes exist to detoxify it. At some threshold level, believed to average around 300–350 ppm, the oxidative enzymes become overwhelmed. Many personal safety gas detectors, such as those used by utility, sewage and petrochemical workers, are set to alarm at as low as 5 to 10 ppm and to go into high alarm at 15 ppm. Detoxification is effected by oxidation to sulfate, which is harmless.[43] Hence, low levels of hydrogen sulfide may be tolerated indefinitely.

Diagnostic of extreme poisoning by H
2
S
is the discolouration of copper coins in the pockets of the victim. Treatment involves immediate inhalation of amyl nitrite, injections of sodium nitrite, or administration of 4-dimethylaminophenol in combination with inhalation of pure oxygen, administration of bronchodilators to overcome eventual bronchospasm, and in some cases hyperbaric oxygen therapy (HBOT).[42] HBOT has clinical and anecdotal support.[44][45][46]

Exposure to lower concentrations can result in eye irritation, a sore throat and cough, nausea, shortness of breath, and fluid in the lungs (pulmonary edema).[42] These effects are believed to be due to the fact that hydrogen sulfide combines with alkali present in moist surface tissues to form sodium sulfide, a caustic.[47] These symptoms usually go away in a few weeks.

Long-term, low-level exposure may result in fatigue, loss of appetite, headaches, irritability, poor memory, and dizziness. Chronic exposure to low level H
2
S
(around 2 ppm) has been implicated in increased miscarriage and reproductive health issues among Russian and Finnish wood pulp workers,[48] but the reports have not (as of circa 1995) been replicated.

Short-term, high-level exposure can induce immediate collapse, with loss of breathing and a high probability of death. If death does not occur, high exposure to hydrogen sulfide can lead to cortical pseudolaminar necrosis, degeneration of the basal ganglia and cerebral edema.[42] Although respiratory paralysis may be immediate, it can also be delayed up to 72 hours.[49]

  • 0.00047 ppm or 0.47 ppb is the odor threshold, the point at which 50% of a human panel can detect the presence of an odor without being able to identify it.[50]
  • 10 ppm is the OSHA permissible exposure limit (PEL) (8 hour time-weighted average).[31]
  • 10–20 ppm is the borderline concentration for eye irritation.
  • 20 ppm is the acceptable ceiling concentration established by OSHA.[31]
  • 50 ppm is the acceptable maximum peak above the ceiling concentration for an 8-hour shift, with a maximum duration of 10 minutes.[31]
  • 50–100 ppm leads to eye damage.
  • At 100–150 ppm the olfactory nerve is paralyzed after a few inhalations, and the sense of smell disappears, often together with awareness of danger.[51][52]
  • 320–530 ppm leads to pulmonary edema with the possibility of death.[42]
  • 530–1000 ppm causes strong stimulation of the central nervous system and rapid breathing, leading to loss of breathing.
  • 800 ppm is the lethal concentration for 50% of humans for 5 minutes' exposure (LC50).
  • Concentrations over 1000 ppm cause immediate collapse with loss of breathing, even after inhalation of a single breath.

Incidents

Hydrogen sulfide was used by the British Army as a chemical weapon during World War I. It was not considered to be an ideal war gas, but, while other gases were in short supply, it was used on two occasions in 1916.[53]

In 1975, a hydrogen sulfide release from an oil drilling operation in Denver City, Texas, killed nine people and caused the state legislature to focus on the deadly hazards of the gas. State Representative E L Short took the lead in endorsing an investigation by the Texas Railroad Commission and urged that residents be warned "by knocking on doors if necessary" of the imminent danger stemming from the gas. A person may die from the second inhalation of the gas, and a warning itself may be too late.[54]

On September 2, 2005, a leak in the propeller room of a Royal Caribbean Cruise Liner docked in Los Angeles resulted in the deaths of 3 crewmen due to a sewage line leak. As a result, all such compartments are now required to have a ventilation system.[55][56]

A dump of toxic waste containing hydrogen sulfide is believed to have caused 17 deaths and thousands of illnesses in Abidjan, on the West African coast, in the 2006 Côte d'Ivoire toxic waste dump.

In September 2008 3 workers were killed and 2 suffered serious injury including long term brain damage at a mushroom growing company in Langley, British Columbia. A valve to a pipe that carried chicken manure, straw and gypsum to the compost fuel for the mushroom growing operation became clogged, and as workers unclogged the valve in a confined space without proper ventilation the hydrogen sulfide that had built up due to anaerobic decomposition of the material was released, positioning the workers in the surrounding area.[57] Investigator said there could have been more fatalities if the pipe had been fully cleared and/or if the wind had changed directions.[58]

In 2014, levels of hydrogen sulfide as high as 83 ppm were detected at a recently built mall in Thailand called Siam Square One at the Siam Square area. Shop tenants at the mall reported health complications such as sinus inflammation, breathing difficulties and eye irritation. After investigation it was determined that the large amount of gas originated from imperfect treatment and disposal of waste water in the building.[59]

In November 2014, a substantial amount of hydrogen sulfide gas shrouded the central, eastern and southeastern parts of Moscow. Residents living in the area were urged to stay indoors by the emergencies ministry. Although the exact source of the gas was not known, blame had been placed on a Moscow oil refinery.[60]

In June 2016, a mother and her daughter were found deceased in their still-running 2006 Porsche Cayenne SUV against a guardrail on Florida's Turnpike, initially thought to be victims of carbon monoxide poisoning.[61][62] Their deaths remained unexplained as the medical examiner waited for results of toxicology tests on the victims,[63] until urine tests revealed that hydrogen sulfide was the cause of death.[64] A report from the Orange-Osceola Medical Examiner's Office indicated that toxic fumes came from the Porsche's starter battery, located under the front passenger seat.[65][66]

In January 2017, three utility workers in Key Largo, Florida, died one by one within seconds of descending into a narrow space beneath a manhole cover to check a section of paved street.[67] In an attempt to save the men, a firefighter who entered the hole without his air tank (because he could not fit through the hole with it) collapsed within seconds and had to be rescued by a colleague.[68][69] The firefighter was airlifted to Jackson Memorial Hospital and later recovered.[70][71] A Monroe County Sheriff officer initially determined that the space contained hydrogen sulfide and methane gas produced by decomposing vegetation.[72]

Suicides

The gas, produced by mixing certain household ingredients, was used in a suicide wave in 2008 in Japan.[73] The wave prompted staff at Tokyo's suicide prevention center to set up a special hot line during "Golden Week", as they received an increase in calls from people wanting to kill themselves during the annual May holiday.[74]

As of 2010, this phenomenon has occurred in a number of US cities, prompting warnings to those arriving at the site of the suicide.[75][76][77][78][79] These first responders, such as emergency services workers or family members are at risk of death from inhaling lethal quantities of the gas, or by fire.[80][81] Local governments have also initiated campaigns to prevent such suicides.

Hydrogen sulfide in the natural environment

Microbial: The sulfur cycle

Sludge from a pond; the black color is due to metal sulfides

Hydrogen sulfide is a central participant in the sulfur cycle, the biogeochemical cycle of sulfur on Earth.[82]

In the absence of oxygen, sulfur-reducing and sulfate-reducing bacteria derive energy from oxidizing hydrogen or organic molecules by reducing elemental sulfur or sulfate to hydrogen sulfide. Other bacteria liberate hydrogen sulfide from sulfur-containing amino acids; this gives rise to the odor of rotten eggs and contributes to the odor of flatulence.

As organic matter decays under low-oxygen (or hypoxic) conditions (such as in swamps, eutrophic lakes or dead zones of oceans), sulfate-reducing bacteria will use the sulfates present in the water to oxidize the organic matter, producing hydrogen sulfide as waste. Some of the hydrogen sulfide will react with metal ions in the water to produce metal sulfides, which are not water-soluble. These metal sulfides, such as ferrous sulfide FeS, are often black or brown, leading to the dark color of sludge.

Several groups of bacteria can use hydrogen sulfide as fuel, oxidizing it to elemental sulfur or to sulfate by using dissolved oxygen, metal oxides (e.g., Fe oxyhydroxides and Mn oxides), or nitrate as electron acceptors.[83]

The purple sulfur bacteria and the green sulfur bacteria use hydrogen sulfide as an electron donor in photosynthesis, thereby producing elemental sulfur. This mode of photosynthesis is older than the mode of cyanobacteria, algae, and plants, which uses water as electron donor and liberates oxygen.

The biochemistry of hydrogen sulfide is a key part of the chemistry of the iron-sulfur world. In this model of the origin of life on Earth, geologically produced hydrogen sulfide is postulated as an electron donor driving the reduction of carbon dioxide.[84]

Animals

Hydrogen sulfide is lethal to most animals, but a few highly specialized species (extremophiles) do thrive in habitats that are rich in this compound.[85]

In the deep sea, hydrothermal vents and cold seeps with high levels of hydrogen sulfide are home to a number of extremely specialized lifeforms, ranging from bacteria to fish.[86] Because of the absence of light at these depths, these ecosystems rely on chemosynthesis rather than photosynthesis.[87]

Freshwater springs rich in hydrogen sulfide are mainly home to invertebrates, but also include a small number of fish: Cyprinodon bobmilleri (a pupfish from Mexico), Limia sulphurophila (a poeciliid from the Dominican Republic), Gambusia eurystoma (a poeciliid from Mexico), and a few Poecilia (poeciliids from Mexico).[85][88] Invertebrates and microorganisms in some cave systems, such as Movile Cave, are adapted to high levels of hydrogen sulfide.[89]

Interstellar and planetary occurrence

Hydrogen sulfide has often been detected in the interstellar medium.[90] It also occurs in the clouds of planets in our solar system.[91][92]

Mass extinctions

A hydrogen sulfide bloom (green) stretching for about 150km along the coast of Namibia. As oxygen-poor water reaches the coast, bacteria in the organic-matter rich sediment produce hydrogen sulfide which is toxic to fish.

Hydrogen sulfide has been implicated in several mass extinctions that have occurred in the Earth's past. In particular, a buildup of hydrogen sulfide in the atmosphere may have caused the Permian-Triassic extinction event 252 million years ago.[93]

Organic residues from these extinction boundaries indicate that the oceans were anoxic (oxygen-depleted) and had species of shallow plankton that metabolized H
2
S
. The formation of H
2
S
may have been initiated by massive volcanic eruptions, which emitted carbon dioxide and methane into the atmosphere, which warmed the oceans, lowering their capacity to absorb oxygen that would otherwise oxidize H
2
S
. The increased levels of hydrogen sulfide could have killed oxygen-generating plants as well as depleted the ozone layer, causing further stress. Small H
2
S
blooms have been detected in modern times in the Dead Sea and in the Atlantic ocean off the coast of Namibia.[93]

See also

  • Amine gas treating
  • Gaseous signaling molecules, also known as gasotransmitters
  • Hydrogen chalcogenide
  • Jenkem
  • Sewer gas
  • Targeted temperature management, also known as induced hypothermia

References

  1. "Hydrogen Sulfide - PubChem Public Chemical Database". The PubChem Project. USA: National Center for Biotechnology Information.
  2. Perrin, D.D. (1982). Ionisation Constants of Inorganic Acids and Bases in Aqueous Solution (2nd ed.). Oxford: Pergamon Press.
  3. Bruckenstein, S.; Kolthoff, I.M., in Kolthoff, I.M.; Elving, P.J. Treatise on Analytical Chemistry, Vol. 1, pt. 1; Wiley, NY, 1959, pp. 432–433.
  4. Patnaik, Pradyot (2002). Handbook of Inorganic Chemicals. McGraw-Hill. ISBN 978-0-07-049439-8.
  5. Zumdahl, Steven S. (2009). Chemical Principles (6th ed.). Houghton Mifflin Company. p. A23. ISBN 978-0-618-94690-7.
  6. "Hydrogen sulfide". npi.gov.au.
  7. "Hydrogen sulfide". Immediately Dangerous to Life and Health Concentrations (IDLH). National Institute for Occupational Safety and Health (NIOSH).
  8. NIOSH Pocket Guide to Chemical Hazards. "#0337". National Institute for Occupational Safety and Health (NIOSH).
  9. Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN 978-0-08-037941-8.
  10. "Hydrogen Sulphide In Well Water". Retrieved 4 September 2018.
  11. Bos, E. M; Van Goor, H; Joles, J. A; Whiteman, M; Leuvenink, H. G (2015). "Hydrogen sulfide: Physiological properties and therapeutic potential in ischaemia". British Journal of Pharmacology. 172 (6): 1479–1493. doi:10.1111/bph.12869. PMC 4369258. PMID 25091411.
  12. May, P.M.; Batka, D.; Hefter, G.; Könignberger, E.; Rowland, D. (2018). "Goodbye to S2-". Chem. Comm. 54 (16): 1980–1983. doi:10.1039/c8cc00187a. PMID 29404555.
  13. Drozdov, A.; Eremets, M. I.; Troyan, I. A. (2014). "Conventional superconductivity at 190 K at high pressures". arXiv:1412.0460 [cond-mat.supr-con].
  14. Cartlidge, Edwin (18 August 2015). "Superconductivity record sparks wave of follow-up physics". Nature. 524 (7565): 277. Bibcode:2015Natur.524..277C. doi:10.1038/nature.2015.18191. PMID 26289188.
  15. Francois Pouliquen; Claude Blanc; Emmanuel Arretz; Ives Labat; Jacques Tournier-Lasserve; Alain Ladousse; Jean Nougayrede; Gérard Savin; Raoul Ivaldi; Monique Nicolas; Jean Fialaire; René Millischer; Charles Azema; Lucien Espagno; Henri Hemmer; Jacques Perrot (200). "Hydrogen Sulfide". Ullmann's Encyclopedia of Chemical Industry. doi:10.1002/14356007.a13_467. ISBN 978-3527306732.
  16. McPherson, William (1913). Laboratory manual. Boston: Ginn and Company. p. 445.
  17. "Why Does My Water Smell Like Rotten Eggs? Hydrogen Sulfide and Sulfur Bacteria in Well Water". Minnesota Department of Health. Minnesota Department of Health. Retrieved 1 December 2014.
  18. T., Hancock, John (2017). Cell signalling (Fourth ed.). Oxford, United Kingdom. ISBN 9780199658480. OCLC 947925636.
  19. Huang, Caleb Weihao; Moore, Philip Keith (2015), "H2S Synthesizing Enzymes: Biochemistry and Molecular Aspects", Chemistry, Biochemistry and Pharmacology of Hydrogen Sulfide, Springer International Publishing, 230, pp. 3–25, doi:10.1007/978-3-319-18144-8_1, ISBN 9783319181431, PMID 26162827
  20. Kabil, Omer; Banerjee, Ruma (2014-02-10). "Enzymology of H2S Biogenesis, Decay and Signaling". Antioxidants & Redox Signaling. 20 (5): 770–782. doi:10.1089/ars.2013.5339. ISSN 1523-0864. PMC 3910450. PMID 23600844.
  21. Kabil, Omer; Vitvitsky, Victor; Xie, Peter; Banerjee, Ruma (2011-07-15). "The Quantitative Significance of the Transsulfuration Enzymes for H2S Production in Murine Tissues". Antioxidants & Redox Signaling. 15 (2): 363–372. doi:10.1089/ars.2010.3781. ISSN 1523-0864. PMC 3118817. PMID 21254839.
  22. Szabo, Csaba (March 2018). "A timeline of hydrogen sulfide (H2S) research: from environmental toxin to biological mediator". Biochemical Pharmacology. 149: 5–19. doi:10.1016/j.bcp.2017.09.010. ISSN 0006-2952. PMC 5862769. PMID 28947277.
  23. Szabo, Csaba; Papapetropoulos, Andreas (October 2017). "International Union of Basic and Clinical Pharmacology. CII: Pharmacological Modulation of H2S Levels: H2S Donors and H2S Biosynthesis Inhibitors". Pharmacological Reviews. 69 (4): 497–564. doi:10.1124/pr.117.014050. ISSN 1521-0081. PMC 5629631. PMID 28978633.
  24. Wang, Rui (April 2012). "Physiological implications of hydrogen sulfide: a whiff exploration that blossomed". Physiological Reviews. 92 (2): 791–896. doi:10.1152/physrev.00017.2011. ISSN 1522-1210. PMID 22535897.
  25. Li, Zhen; Polhemus, David J.; Lefer, David J. (17 August 2018). "Evolution of Hydrogen Sulfide Therapeutics to Treat Cardiovascular Disease". Circulation Research. 123 (5): 590–600. doi:10.1161/CIRCRESAHA.118.311134. ISSN 1524-4571. PMID 30355137.
  26. Kimura, Hideo (2019-01-18). "Signalling by hydrogen sulfide and polysulfides via protein S-sulfuration". British Journal of Pharmacology. doi:10.1111/bph.14579. ISSN 1476-5381. PMID 30657595.
  27. Khazaei, A.; et al. (2012). "Synthesis of thiophenols using sodium sulfide in acidic media". Synlett. 23 (13): 1893–1896. doi:10.1055/s-0032-1316557.
  28. Stampler, Laura. "A Stinky Compound May Protect Against Cell Damage, Study Finds". Time. Time. Retrieved 1 December 2014.
  29. Aslami, H; Schultz, MJ; Juffermans, NP (2009). "Potential applications of hydrogen sulfide-induced suspended animation". Current Medicinal Chemistry. 16 (10): 1295–303. doi:10.2174/092986709787846631. PMID 19355886.
  30. "Hydrogen Sulfide". Earthworks. Retrieved 2020-04-18.
  31. Agency for Toxic Substances and Disease Registry (July 2006). "Toxicological Profile For Hydrogen Sulfide" (PDF). p. 154. Retrieved 2012-06-20.
  32. OnePetro. "Home - OnePetro". onepetro.org.
  33. "Hydrogen Sulfide" (PDF). Agency for Toxic Substances and Disease Registry. December 2016.
  34. Jang, Yong-Chul; Townsend, Timothy (2001). "Sulfate leaching from recovered construction and demolition debris fines". Advances in Environmental Research. 5 (3): 203–217. doi:10.1016/S1093-0191(00)00056-3.
  35. Cavinato, C (2013) [2013]. "Anaerobic digestion fundamentals" (PDF).
  36. "Sulfur-oxidizing Bacteria in Environmental Technology". ResearchGate. Retrieved 2019-03-26.
  37. Lemley, Ann T.; Schwartz, John J.; Wagenet, Linda P. "Hydrogen Sulfide in Household Drinking Water" (PDF). Cornell University.
  38. "Hydrogen Sulfide (Rotten Egg Odor) in Pennsylvania Groundwater Wells". Penn State. Penn State College of Agricultural Sciences. Retrieved 1 December 2014.
  39. McFarland, Mark L.; Provin, T. L. "Hydrogen Sulfide in Drinking Water Treatment Causes and Alternatives" (PDF). Texas A&M University. Retrieved 1 December 2014.
  40. "Why Does My Water Smell Like Rotten Eggs?". Minnesota Department of Health. Retrieved 20 January 2020.
  41. Iowa State University, Department of Chemistry MSDS. "Hydrogen Sulfide Material Safety Data Sheet" (PDF). Archived from the original (PDF) on 2009-03-27. Retrieved 2009-03-14.
  42. Lindenmann, J.; Matzi, V.; Neuboeck, N.; Ratzenhofer-Komenda, B.; Maier, A; Smolle-Juettner, F. M. (December 2010). "Severe hydrogen sulphide poisoning treated with 4-dimethylaminophenol and hyperbaric oxygen". Diving and Hyperbaric Medicine. 40 (4): 213–217. PMID 23111938. Retrieved 2013-06-07.
  43. Ramasamy, S.; Singh, S.; Taniere, P.; Langman, M. J. S.; Eggo, M. C. (2006). "Sulfide-detoxifying enzymes in the human colon are decreased in cancer and upregulated in differentiation". Am. J. Physiol. Gastrointest. Liver Physiol. 291 (2): G288–96. doi:10.1152/ajpgi.00324.2005. PMID 16500920.
  44. Gerasimon, G.; Bennett, S.; Musser, J.; Rinard, J. (May 2007). "Acute hydrogen sulfide poisoning in a dairy farmer". Clin. Toxicol. 45 (4): 420–423. doi:10.1080/15563650601118010. PMID 17486486.
  45. Belley, R.; Bernard, N.; Côté, M; Paquet, F.; Poitras, J. (July 2005). "Hyperbaric oxygen therapy in the management of two cases of hydrogen sulfide toxicity from liquid manure". CJEM. 7 (4): 257–261. doi:10.1017/s1481803500014408. PMID 17355683. Archived from the original on 2010-09-11. Retrieved 2008-07-22.
  46. Hsu, P.; Li, H.-W.; Lin, Y.-T. (1987). "Acute hydrogen sulfide poisoning treated with hyperbaric oxygen". J. Hyperbaric Med. 2 (4): 215–221. ISSN 0884-1225. Retrieved 2008-07-22.
  47. Lewis, R.J. (1996). Sax's Dangerous Properties of Industrial Materials. 1–3 (9th ed.). New York, NY: Van Nostrand Reinhold.
  48. Hemminki, K.; Niemi, M. L. (1982). "Community study of spontaneous abortions: relation to occupation and air pollution by sulfur dioxide, hydrogen sulfide, and carbon disulfide". Int. Arch. Occup. Environ. Health. 51 (1): 55–63. doi:10.1007/bf00378410. PMID 7152702.
  49. "The chemical suicide phenomenon". Firerescue1.com. 2011-02-07. Retrieved 2013-12-19.
  50. Iowa State University Extension (May 2004). "The Science of Smell Part 1: Odor perception and physiological response" (PDF). PM 1963a. Retrieved 2012-06-20.
  51. USEPA; Health and Environmental Effects Profile for Hydrogen Sulfide p.118-8 (1980) ECAO-CIN-026A
  52. Zenz, C.; Dickerson, O.B.; Horvath, E.P. (1994). Occupational Medicine (3rd ed.). St. Louis, MO. p. 886.
  53. Foulkes, Charles Howard (2001) [First published Blackwood & Sons, 1934]. "Gas!" The story of the special brigade. Published by Naval & Military P. p. 105. ISBN 978-1-84342-088-0.
  54. Howard Swindle (June 1975). "The Deadly Smell of Success". Texas Monthly. Retrieved December 14, 2010 via Google Books.
  55. "LA County Department of Public Health" (PDF). County of Los Angeles: Department of Public Health. Archived from the original (PDF) on 2017-02-18. Retrieved 2017-06-11.
  56. Becerra, Hector; Pierson, David (2005-09-03). "Gas Kills 3 Crewmen on Ship". Los Angeles Times.
  57. Ferguson, Dan (September 16, 2011). "Details of Langley mushroom farm tragedy finally disclosed". Abbotsford News. Retrieved April 13, 2020.
  58. Theodore, Terri (May 8, 2012). "Dozens could have died because of owner's negligence in B.C. mushroom farm incident: investigator". The Canadian Press. The Globe and Mail. Retrieved April 13, 2020.
  59. "Do not breathe: Dangerous, toxic gas found at Siam Square One". Coconuts Bangkok. Coconuts Media. 2014-10-21. Retrieved 20 November 2014.
  60. "Russian capital Moscow shrouded in noxious gas". BBC News. British Broadcasting Corporation. 2014-11-10. Retrieved 1 December 2014.
  61. "Sources: Mom, daughter found dead in Porsche likely died from carbon monoxide". WFTV. 7 June 2016. Retrieved 28 April 2017. Both had red skin and rash-like symptoms, and had vomited, sources said.
  62. Salinger, Tobias (4 October 2016). "Woman, girl died after inhaling hydrogen sulfide, coroners say". New York Daily News. Retrieved 28 April 2017.
  63. Lotan, Gal Tziperman (4 October 2016). "Hydrogen sulfide inhalation killed mother, toddler found on Florida's Turnpike in June". Orlando Sentinel. Retrieved 28 April 2017.
  64. Zilber, Ariel (2016-10-05). "Florida woman and her daughter who died in Porsche inhaled toxic gas". dailymail.co.uk. Mail Online. Retrieved 28 April 2017.
  65. Kealing, Bob. "Medical examiner confirms suspected cause of deaths in Turnpike mystery". Archived from the original on 2016-10-05. Retrieved 2016-10-04.
  66. Bell, Lisa (19 March 2017). "Hidden car dangers you should be aware of". ClickOrlando.com. Produced by Donovan Myrie. WKMG-TV. Retrieved 28 April 2017. Porsche Cayennes, along with a few other vehicles, have their batteries in the passenger compartment.
  67. "One by one, 3 utility workers descended into a manhole. One by one, they died". www.washingtonpost.com.
  68. Herrin, Becky (16 January 2017). "Detectives investigating deaths of three men". floridakeyssheriff.blogspot.com. Monroe County Sheriff's Office. Retrieved 28 April 2017.
  69. Goodhue, David (17 January 2017). "Firefighter who tried to save 3 men in a manhole is fighting for his life". Miami Herald. Retrieved 28 April 2017.
  70. "Key Largo firefighter takes first steps since nearly getting killed". WSVN. 18 January 2017. Retrieved 28 April 2017.
  71. "Firefighter who survived Key Largo rescue attempt that killed 3 leaves hospital". Sun-Sentinel. Associated Press. 26 January 2017. Retrieved 28 April 2017.
  72. Rabin, Charles; Goodhue, David (16 January 2017). "Three Keys utility workers die in wastewater trench". Miami Herald. Retrieved 28 April 2017.
  73. "Dangerous Japanese 'Detergent Suicide' Technique Creeps Into U.S". Wired.com. March 13, 2009.
  74. Namiki, Noriko (2008-05-22). "Terrible Twist in Japan Suicide Spates - ABC News". Abcnews.go.com. Retrieved 2013-12-19.
  75. http://info.publicintelligence.net/LARTTAChydrogensulfide.pdf
  76. http://info.publicintelligence.net/MAchemicalsuicide.pdf
  77. http://info.publicintelligence.net/illinoisH2Ssuicide.pdf
  78. http://info.publicintelligence.net/NYhydrogensulfide.pdf
  79. http://info.publicintelligence.net/KCTEWhydrogensulfide.pdf
  80. dhmh.maryland.gov Archived January 3, 2012, at the Wayback Machine
  81. Scoville, Dean (April 2011). "Chemical Suicides - Article - POLICE Magazine". Policemag.com. Retrieved 2013-12-19.
  82. Barton, Larry L.; Fardeau, Marie-Laure; Fauque, Guy D. (2014). "Chapter 10. Hydrogen Sulfide: A Toxic Gas Produced by Dissimilatory Sulfate and Sulfur Reduction and Consumed by Microbial Oxidation". In Kroneck, Peter M.H.; Sosa Torres, Martha E. (eds.). The Metal-Driven Biogeochemistry of Gaseous Compounds in the Environment. Metal Ions in Life Sciences. 14. Springer. pp. 237–277. doi:10.1007/978-94-017-9269-1_10. ISBN 978-94-017-9268-4. PMID 25416397.
  83. Jørgensen, B. B.; Nelson, D. C. (2004). "Sulfide oxidation in marine sediments: Geochemistry meets microbiology". In Amend, J. P.; Edwards, K. J.; Lyons, T. W. (eds.). Sulfur Biogeochemistry – Past and Present. Geological Society of America. pp. 36–81.
  84. "The very first surface organism can be characterized as a catalyst for accelerating the formation of pyrite by providing a catalytic pathway for the flow of electrons from hydrogen sulfide to carbon dioxide." Wächtershäuser, Günter (1988-12-01). "Before enzymes and templates: theory of surface metabolism". Microbiol. Mol. Biol. Rev. 52 (4): 452–84. doi:10.1128/MMBR.52.4.452-484.1988. PMC 373159. PMID 3070320. Retrieved July 25, 2015.
  85. Tobler, M; Riesch, R.; García de León, F. J.; Schlupp, I.; Plath, M. (2008). "Two endemic and endangered fishes, Poecilia sulphuraria (Álvarez, 1948) and Gambusia eurystoma Miller, 1975 (Poeciliidae, Teleostei) as only survivors in a small sulphidic habitat". Journal of Fish Biology. 72 (3): 523–533. doi:10.1111/j.1095-8649.2007.01716.x.
  86. Bernardino, Angelo F.; Levin, Lisa A.; Thurber, Andrew R.; Smith, Craig R. (2012). "Comparative Composition, Diversity and Trophic Ecology of Sediment Macrofauna at Vents, Seeps and Organic Falls". PLoS ONE. 7 (4): e33515. Bibcode:2012PLoSO...733515B. doi:10.1371/journal.pone.0033515. PMC 3319539. PMID 22496753.
  87. "Hydrothermal Vents". Marine Society of Australia. Retrieved 28 December 2014.
  88. Palacios, Maura; Arias-Rodríguez, Lenín; Plath, Martin; Eifert, Constanze; Lerp, Hannes; Lamboj, Anton; Voelker, Gary; Tobler, Michael (2013). "The Rediscovery of a Long Described Species Reveals Additional Complexity in Speciation Patterns of Poeciliid Fishes in Sulfide Springs". PLoS ONE. 8 (8): e71069. Bibcode:2013PLoSO...871069P. doi:10.1371/journal.pone.0071069. PMC 3745397. PMID 23976979.
  89. Kumaresan, Deepak; Wischer, Daniela; Stephenson, Jason; Hillebrand-Voiculescu, Alexandra; Murrell, J. Colin (2014). "Microbiology of Movile Cave—A Chemolithoautotrophic Ecosystem". Geomicrobiology Journal. 31 (3): 186–193. doi:10.1080/01490451.2013.839764. ISSN 0149-0451.
  90. Despois, D. (1999). "Radio line observations of molecular and isotopic species in comet C/1995 O1 (Hale-Bopp) Implications on the interstellar origin of cometary ices". Earth, Moon, and Planets. 79: 103–124. doi:10.1023/A:1006229131864.
  91. Irwin, P. G. J.; Toledo, D.; Garland, R.; Teanby, N. A.; Fletcher, L. N.; Orton, G. A.; Bézard, B. (2018). "Detection of hydrogen sulfide above the clouds in Uranus's atmosphere". Nature Astronomy. 2 (5): 420–427. Bibcode:2018NatAs...2..420I. doi:10.1038/s41550-018-0432-1. hdl:2381/42547.CS1 maint: uses authors parameter (link)
  92. Lissauer, Jack J.; de Pater, Imke (2019). Fundamental Planetary Sciences : physics, chemistry, and habitability. New York, NY, USA: Cambridge University Press. pp. 149–152. ISBN 9781108411981.
  93. "Impact from the Deep". Scientific American. October 2006.

Additional resources

This article is issued from Wikipedia. The text is licensed under Creative Commons - Attribution - Sharealike. Additional terms may apply for the media files.