Hydrogen sulfide

Last updated
Hydrogen sulfide
Hydrogen-sulfide-2D-dimensions.svg
Ball-and-stick model of hydrogen sulfide Hydrogen-sulfide-3D-balls.png
Ball-and-stick model of hydrogen sulfide
Spacefill model of hydrogen sulfide Hydrogen-sulfide-3D-vdW.svg
Spacefill model of hydrogen sulfide
  Sulfur, S
  Hydrogen, H
Names
Systematic IUPAC name
Hydrogen sulfide [1]
Other names
  • Dihydrogen monosulfide
  • Sour gas
  • Dihydrogen sulfide
  • Sewer gas
  • Egg gas
  • Sulfane
  • Sulfurated hydrogen
  • Sulfureted hydrogen
  • Sulfuretted hydrogen
  • Sulfur hydride
  • Hydrosulfuric acid
  • Hydrothionic acid
  • Thiohydroxic acid
  • Sulfhydric acid
Identifiers
3D model (JSmol)
3DMet
3535004
ChEBI
ChEMBL
ChemSpider
ECHA InfoCard 100.029.070 OOjs UI icon edit-ltr-progressive.svg
EC Number
  • 231-977-3
303
KEGG
MeSH Hydrogen+sulfide
PubChem CID
RTECS number
  • MX1225000
UNII
UN number 1053
  • InChI=1S/H2S/h1H2 Yes check.svgY
    Key: RWSOTUBLDIXVET-UHFFFAOYSA-N Yes check.svgY
  • InChI=1/H2S/h1H2
    Key: RWSOTUBLDIXVET-UHFFFAOYAJ
Properties
H2S
Molar mass 34.08 g·mol−1
AppearanceColorless gas
Odor Foul, pungent, like that of rotten eggs
Density 1.539 g.L−1 (0°C) [2]
Melting point −85.5 [3]  °C (−121.9 °F; 187.7 K)
Boiling point −59.55 [3]  °C (−75.19 °F; 213.60 K)
3.980 g dm−3 (at 20 °C) [4]
Vapor pressure 1740 kPa (at 21 °C)
Acidity (pKa)7.0 [5] [6]
Conjugate acid Sulfonium
Conjugate base Bisulfide
25.5·10−6 cm3/mol
1.000644 (0 °C) [2]
Structure
C2v
Bent
0.97 D
Thermochemistry
1.003 J K−1 g−1
Std molar
entropy
(S298)
206 J mol−1 K−1 [7]
−21 kJ mol−1 [7]
Hazards
Occupational safety and health (OHS/OSH):
Main hazards
Flammable and highly toxic
GHS labelling:
GHS-pictogram-flamme.svg GHS-pictogram-skull.svg GHS-pictogram-pollu.svg
Danger
H220, H330, H400
P210, P260, P271, P273, P284, P304+P340, P310, P320, P377, P381, P391, P403, P403+P233, P405, P501
NFPA 704 (fire diamond)
NFPA 704.svgHealth 4: Very short exposure could cause death or major residual injury. E.g. VX gasFlammability 4: Will rapidly or completely vaporize at normal atmospheric pressure and temperature, or is readily dispersed in air and will burn readily. Flash point below 23 °C (73 °F). E.g. propaneInstability 0: Normally stable, even under fire exposure conditions, and is not reactive with water. E.g. liquid nitrogenSpecial hazards (white): no code
4
4
0
Flash point −82.4 °C (−116.3 °F; 190.8 K) [8]
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) [9]
  • 600 ppm (human, 30 min)
  • 800 ppm (human, 5 min) [9]
NIOSH (US health exposure limits):
PEL (Permissible)
C 20 ppm; 50 ppm [10-minute maximum peak] [10]
REL (Recommended)
C 10 ppm (15 mg/m3) [10-minute] [10]
IDLH (Immediate danger)
100 ppm [10]
Related compounds
Related hydrogen chalcogenides
Related compounds
Phosphine
Except where otherwise noted, data are given for materials in their standard state (at 25 °C [77 °F], 100 kPa).
X mark.svgN  verify  (what is  Yes check.svgYX mark.svgN ?)

Hydrogen sulfide is a chemical compound with the formula H 2 S . It is a colorless chalcogen-hydride gas, and is poisonous, corrosive, and flammable, with trace amounts in ambient atmosphere having a characteristic foul odor of rotten eggs. [11] Swedish chemist Carl Wilhelm Scheele is credited with having discovered the chemical composition of purified hydrogen sulfide in 1777. [12]

Contents

Hydrogen sulfide is toxic to humans and most other animals by inhibiting cellular respiration in a manner similar to hydrogen cyanide. When it is inhaled or its salts are ingested in high amounts,[ clarification needed ] damage to organs occurs rapidly with symptoms ranging from breathing difficulties to convulsions and death. [13] [14] Despite this, the human body produces small amounts of this sulfide and its mineral salts, and uses it as a signalling molecule. [15]

Hydrogen sulfide is often produced from the microbial breakdown of organic matter in the absence of oxygen, such as in swamps and sewers; this process is commonly known as anaerobic digestion, which is done by sulfate-reducing microorganisms. It also occurs in volcanic gases, natural gas deposits, and sometimes in well-drawn water.

Properties

Hydrogen sulfide is slightly denser than air. A mixture of H2S and air can be explosive.

Oxidation

In general, hydrogen sulfide acts as a reducing agent, as indicated by its ability to reduce sulfur dioxide in the Claus process. Hydrogen sulfide burns in oxygen with a blue flame to form sulfur dioxide (SO2) and water:

2 H2S + 3 O2 → 2 SO2 + 2 H2O

If an excess of oxygen is present, sulfur trioxide (SO3) is formed, which quickly hydrates to sulfuric acid:

H2S + 2 O2 → H2SO4

Acid-base properties

Hydrogen sulfide is a weak acid, forming SH.

It 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. [16]

Extreme temperatures and pressures

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. [17] 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. [18]

Hydrogen sulfide decomposes without a presence of a catalyst under atmospheric pressure around 1200 °C into hydrogen and sulfur. [19]

Tarnishing

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 or electrolysis often liberates hydrogen sulfide. Hydrogen sulfide is also responsible for tarnishing on various metals including copper and silver; the chemical responsible for black toning found on silver coins is silver sulfide (Ag2S), which is produced when the silver on the surface of the coin reacts with atmospheric hydrogen sulfide. [20] Coins that have been subject to toning by hydrogen sulfide and other sulfur-containing compounds may have the toning add to the numismatic value of a coin based on aesthetics, as the toning may produce thin-film interference, resulting in the coin taking on an attractive coloration. [21] Coins can also be intentionally treated with hydrogen sulfide to induce toning, though artificial toning can be distinguished from natural toning, and is generally criticised among collectors. [22]

Production

Hydrogen sulfide is most commonly obtained by its separation from sour gas, which is natural gas with a high content of H2S. 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. [23]

S + H2 → H2S

The very favorable thermodynamics for the hydrogenation of sulfur implies that the dehydrogenation (or cracking of hydrogen sulfide would require very high temperatures. [24]

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 H2S:

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: [25]

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.

Biosynthesis

Hydrogen sulfide can be generated in cells via enzymatic or non-enzymatic pathways. Three enzymes catalyze formation of H
2
S
: cystathionine γ-lyase (CSE), cystathionine β-synthetase (CBS), and 3-mercaptopyruvate sulfurtransferase (3-MST). [26] CBS and CSE are the main proponents of H2S biogenesis, which follows the trans-sulfuration pathway. [27] These enzymes have been identified in a breadth of biological cells and tissues, and their activity is induced by a number of disease states. [28] These enzymes are characterized by the transfer of a sulfur atom from methionine to serine to form a cysteine molecule. [27] 3-MST also contributes to hydrogen sulfide production by way of the cysteine catabolic pathway. [28] [27] 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 derived from proteins such as ferredoxins and Rieske proteins. [28]

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.

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. [29]


Signalling role

H2S in the body acts as a gaseous signaling molecule with implications for health and in diseases. [26] [30] [31]

Hydrogen sulfide is involved in vasodilation in animals, as well as in increasing seed germination and stress responses in plants. [32] Hydrogen sulfide signaling is moderated by reactive oxygen species (ROS) and reactive nitrogen species (RNS). [32] H2S has been shown to interact with NO resulting in several different cellular effects, as well as the formation of another signal called nitrosothiol. [32] Hydrogen sulfide is also known to increase the levels of glutathione, which acts to reduce or disrupt ROS levels in cells. [32]

The field of H2S biology has advanced from environmental toxicology to investigate the roles of endogenously produced H2S in physiological conditions and in various pathophysiological states. [33] H2S has been implicated in cancer and Down syndrome and vascular disease. [34] [35] [36] [37]

It inhibits Complex IV of the mitochondrial electron transport chain, which effectively reduces ATP generation and biochemical activity within cells. [32]

Uses

Production of sulfur

Hydrogen sulfide is mainly consumed as a precursor to elemental sulfur. This conversion, called the Claus process, involves partial oxidation to sulfur dioxide. The latter reacts with hydrogen sulfide to give elemental sulfur. The conversion is catalyzed by alumina. [38]

2H2S + SO2→ 3S + 2H2O

Production of thioorganic compounds

Many fundamental organosulfur compounds are produced using hydrogen sulfide. These include methanethiol, ethanethiol, and thioglycolic acid. [23] Hydrosulfides can be used in the production of thiophenol. [39]

Production of metal sulfides

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

Sodium sulfides are used in the paper making industry. Specifically, salts of SH break bonds between lignin and cellulose components of pulp in the Kraft process. [23]

As indicated above, many metal ions react with hydrogen sulfide to give the corresponding metal sulfides. Oxidic ores are sometimes treated with hydrogen sulfide to give the corresponding metal sulfides which are more readily purified byy flotation. [23] Metal parts are sometimes passivated with hydrogen sulfide. Catalysts used in hydrodesulfurization are routinely activated with hydrogen sulfide.

Hydrogen sulfide was a reagent 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 H2S. The components of the resulting solid are then identified by their reactivity.

Miscellaneous applications

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

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. Oxygen demand was also reduced, thereby protecting against hypoxia. In addition, hydrogen sulfide has been shown to reduce inflammation in various situations. [40]

Occurrence

Deposit of sulfur on a rock, caused by volcanic gas Deposit from hydrogen sulphide.jpg
Deposit of sulfur on a rock, caused by volcanic gas

Volcanoes and some hot springs (as well as cold springs) emit some H2S. Hydrogen sulfide can be present naturally in well water, often as a result of the action of sulfate-reducing bacteria. [41] [ better source needed ] Hydrogen sulfide is produced by the human body in small quantities through bacterial breakdown of proteins containing sulfur in the intestinal tract, therefore it contributes to the characteristic odor of flatulence. It is also produced in the mouth (halitosis). [42]

A portion of global H2S emissions are due to human activity. By far the largest industrial source of H2S is petroleum refineries: The hydrodesulfurization process liberates sulfur from petroleum by the action of hydrogen. The resulting H2S 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. H2S 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 concentrations of H2S were observed in the Bakken formation crude, possibly due to oil field practices, 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". [43]

Besides living near 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. [44]

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 sulfate 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 [45] 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. [46] 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. [47]

Removal from water

A number of processes have been designed to remove hydrogen sulfide from drinking water. [48]

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. [49]
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. [50]
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 ⇌ [RNH3]+ + 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 AmineTreating.png
Process flow diagram of a typical amine treating process used in petroleum refineries, natural gas processing plants and other industrial facilities

Safety

The underground mine gas term for foul-smelling hydrogen sulfide-rich gas mixtures is stinkdamp. Hydrogen sulfide is a highly toxic and flammable gas (flammable range: 4.3–46%). It can poison several systems in the body, although the nervous system is most affected.[ citation needed ] The toxicity of H2S is comparable with that of carbon monoxide. [51] It binds with iron in the mitochondrial cytochrome enzymes, thus preventing cellular respiration. Its toxic properties were described in detail in 1843 by Justus von Liebig. [52]

Even before hydrogen sulfide was discovered, Italian physician Bernardino Ramazzini hypothesized in his 1713 book De Morbis Artificum Diatriba that occupational diseases of sewer-workers and blackening of coins in their clothes may be caused by an unknown invisible volatile acid (moreover, in late 18th century toxic gas emanation from Paris sewers became a problem for the citizens and authorities). [53]

Although very pungent at first (it smells like rotten eggs [54] ), it quickly deadens the sense of smell, creating temporary anosmia, [55] so victims may be unaware of its presence until it is too late. Safe handling procedures are provided by its safety data sheet (SDS). [56]

Low-level exposure

Since hydrogen sulfide occurs naturally in the body, the environment, and the gut, enzymes exist to metabolize 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. Metabolism causes oxidation to sulfate, which is harmless. [57] Hence, low levels of hydrogen sulfide may be tolerated indefinitely.

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). [51] These effects are believed to be due to hydrogen sulfide combining with alkali present in moist surface tissues to form sodium sulfide, a caustic. [58] These symptoms usually subside 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 H2S (around 2 ppm) has been implicated in increased miscarriage and reproductive health issues among Russian and Finnish wood pulp workers, [59] but the reports have not (as of 1995) been replicated.

High-level exposure

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. [51] Although respiratory paralysis may be immediate, it can also be delayed up to 72 hours. [60] Diagnostic of extreme poisoning by H2S is the discolouration of copper coins in the pockets of the victim.

Inhalation of H2S resulted in about 7 workplace deaths per year in the U.S. (2011–2017 data), second only to carbon monoxide (17 deaths per year) for workplace chemical inhalation deaths. [61]

Exposure thresholds

  • Exposure limits stipulated by the United States government: [62]
    • 10 ppm REL-Ceiling (NIOSH): recommended permissible exposure ceiling (the recommended level that must not be exceeded, except once for 10 min. in an 8-hour shift, if no other measurable exposure occurs)
    • 20 ppm PEL-Ceiling (OSHA): permissible exposure ceiling (the level that must not be exceeded, except once for 10 min. in an 8-hour shift, if no other measurable exposure occurs)
    • 50 ppm PEL-Peak (OSHA): peak permissible exposure (the level that must never be exceeded)
    • 100 ppm IDLH (NIOSH): immediately dangerous to life and health (the level that interferes with the ability to escape)
  • 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. [63]
  • 10–20 ppm is the borderline concentration for eye irritation.
  • 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. [64] [65]
  • 320–530 ppm leads to pulmonary edema with the possibility of death. [51]
  • 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.

Treatment

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). [51] HBOT has clinical and anecdotal support. [66] [67] [68]

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, partially due to its flammability and because the distinctive smell could be detected from even a small leak, alerting the enemy to the presence of the gas. It was nevertheless used on two occasions in 1916 when other gases were in short supply. [69]

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. [70] [71]

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, three workers were killed and two 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, poisoning the workers in the surrounding area. [72] Investigator said there could have been more fatalities if the pipe had been fully cleared and/or if the wind had changed directions. [73]

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. [74]

In 2014, hydrogen sulfide gas killed workers at the Promenade shopping center in North Scottsdale, Arizona, USA [75] after climbing into 15ft deep chamber without wearing personal protective gear. "Arriving crews recorded high levels of hydrogen cyanide and hydrogen sulfide coming out of the sewer."

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. [76]

In June 2016, a mother and her daughter were found dead in their still-running 2006 Porsche Cayenne SUV against a guardrail on Florida's Turnpike, initially thought to be victims of carbon monoxide poisoning. [77] [78] Their deaths remained unexplained as the medical examiner waited for results of toxicology tests on the victims, [79] until urine tests revealed that hydrogen sulfide was the cause of death. 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. [80] [81]

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. [82] 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. [83] The firefighter was airlifted to Jackson Memorial Hospital and later recovered. [84] [85] A Monroe County Sheriff officer initially determined that the space contained hydrogen sulfide and methane gas produced by decomposing vegetation. [86]

On May 24, 2018, two workers were killed, another seriously injured, and 14 others hospitalized by hydrogen sulfide inhalation at a Norske Skog paper mill in Albury, New South Wales. [87] [88] An investigation by SafeWork NSW found that the gas was released from a tank used to hold process water. The workers were exposed at the end of a 3-day maintenance period. Hydrogen sulfide had built up in an upstream tank, which had been left stagnant and untreated with biocide during the maintenance period. These conditions allowed sulfate-reducing bacteria to grow in the upstream tank, as the water contained small quantities of wood pulp and fiber. The high rate of pumping from this tank into the tank involved in the incident caused hydrogen sulfide gas to escape from various openings around its top when pumping was resumed at the end of the maintenance period. The area above it was sufficiently enclosed for the gas to pool there, despite not being identified as a confined space by Norske Skog. One of the workers who was killed was exposed while investigating an apparent fluid leak in the tank, while the other who was killed and the worker who was badly injured were attempting to rescue the first after he collapsed on top of it. In a resulting criminal case, Norske Skog was accused of failing to ensure the health and safety of its workforce at the plant to a reasonably practicable extent. It pleaded guilty, and was fined AU$1,012,500 and ordered to fund the production of an anonymized educational video about the incident. [89] [90] [87] [91]

In October 2019, an Odessa, Texas employee of Aghorn Operating Inc. and his wife were killed due to a water pump failure. Produced water with a high concentration of hydrogen sulfide was released by the pump. The worker died while responding to an automated phone call he had received alerting him to a mechanical failure in the pump, while his wife died after driving to the facility to check on him. [92] A CSB investigation cited lax safety practices at the facility, such as an informal lockout-tagout procedure and a nonfunctioning hydrogen sulfide alert system. [93]

Suicides

The gas, produced by mixing certain household ingredients, was used in a suicide wave in 2008 in Japan. [94] The wave prompted staff at Tokyo's suicide prevention center to set up a special hotline during "Golden Week", as they received an increase in calls from people wanting to kill themselves during the annual May holiday. [95]

As of 2010, this phenomenon has occurred in a number of US cities, prompting warnings to those arriving at the site of the suicide. [96] [97] [98] [99] [100] These first responders, such as emergency services workers or family members are at risk of death or injury from inhaling the gas, or by fire. [101] [102] Local governments have also initiated campaigns to prevent such suicides.

In 2020, H2S ingestion was used as a suicide method by Japanese pro wrestler Hana Kimura. [103]

Hydrogen sulfide in the natural environment

Microbial: The sulfur cycle

Sludge from a pond; the black color is due to metal sulfides Teichschlamm1.jpg
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. [104]

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., iron oxyhydroxides and manganese oxides), or nitrate as electron acceptors. [105]

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. [106]

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. [107]

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.[ which? ] [108] Because of the absence of sunlight at these depths, these ecosystems rely on chemosynthesis rather than photosynthesis. [109]

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). [107] [110] Invertebrates and microorganisms in some cave systems, such as Movile Cave, are adapted to high levels of hydrogen sulfide. [111]

Interstellar and planetary occurrence

Hydrogen sulfide has often been detected in the interstellar medium. [112] It also occurs in the clouds of planets in our solar system. [113] [114]

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 Emissions off of Africa.jpg
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, or at least contributed to, the Permian-Triassic extinction event 252 million years ago. [115] [116] [117]

Organic residues from these extinction boundaries indicate that the oceans were anoxic (oxygen-depleted) and had species of shallow plankton that metabolized H2S. The formation of H2S 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 H2S. The increased levels of hydrogen sulfide could have killed oxygen-generating plants as well as depleted the ozone layer, causing further stress. Small H2S blooms have been detected in modern times in the Dead Sea and in the Atlantic ocean off the coast of Namibia. [115]

See also

Related Research Articles

<span class="mw-page-title-main">Sulfur</span> Chemical element, symbol S and atomic number 16

Sulfur (also spelled sulphur in British English) is a chemical element; it has symbol S and atomic number 16. It is abundant, multivalent and nonmetallic. Under normal conditions, sulfur atoms form cyclic octatomic molecules with the chemical formula S8. Elemental sulfur is a bright yellow, crystalline solid at room temperature.

Sulfide (British English also sulphide) is an inorganic anion of sulfur with the chemical formula S2− or a compound containing one or more S2− ions. Solutions of sulfide salts are corrosive. Sulfide also refers to large families of inorganic and organic compounds, e.g. lead sulfide and dimethyl sulfide. Hydrogen sulfide (H2S) and bisulfide (SH) are the conjugate acids of sulfide.

Anaerobic respiration is respiration using electron acceptors other than molecular oxygen (O2). Although oxygen is not the final electron acceptor, the process still uses a respiratory electron transport chain.

The purple sulfur bacteria (PSB) are part of a group of Pseudomonadota capable of photosynthesis, collectively referred to as purple bacteria. They are anaerobic or microaerophilic, and are often found in stratified water environments including hot springs, stagnant water bodies, as well as microbial mats in intertidal zones. Unlike plants, algae, and cyanobacteria, purple sulfur bacteria do not use water as their reducing agent, and therefore do not produce oxygen. Instead, they can use sulfur in the form of sulfide, or thiosulfate (as well, some species can use H2, Fe2+, or NO2) as the electron donor in their photosynthetic pathways. The sulfur is oxidized to produce granules of elemental sulfur. This, in turn, may be oxidized to form sulfuric acid.

<span class="mw-page-title-main">Chromatiaceae</span> Family of purple sulfur bacteria

The Chromatiaceae are one of the two families of purple sulfur bacteria, together with the Ectothiorhodospiraceae. They belong to the order Chromatiales of the class Gammaproteobacteria, which is composed by unicellular Gram-negative organisms. Most of the species are photolithoautotrophs and conduct an anoxygenic photosynthesis, but there are also representatives capable of growing under dark and/or microaerobic conditions as either chemolithoautotrophs or chemoorganoheterotrophs.

<span class="mw-page-title-main">Sulfate-reducing microorganism</span> Microorganisms that "breathe" sulfates

Sulfate-reducing microorganisms (SRM) or sulfate-reducing prokaryotes (SRP) are a group composed of sulfate-reducing bacteria (SRB) and sulfate-reducing archaea (SRA), both of which can perform anaerobic respiration utilizing sulfate (SO2−
4
) as terminal electron acceptor, reducing it to hydrogen sulfide (H2S). Therefore, these sulfidogenic microorganisms "breathe" sulfate rather than molecular oxygen (O2), which is the terminal electron acceptor reduced to water (H2O) in aerobic respiration.

In chemistry, disproportionation, sometimes called dismutation, is a redox reaction in which one compound of intermediate oxidation state converts to two compounds, one of higher and one of lower oxidation states. The reverse of disproportionation, such as when a compound in an intermediate oxidation state is formed from precursors of lower and higher oxidation states, is called comproportionation, also known as synproportionation.

<span class="mw-page-title-main">Iron(II) sulfide</span> Chemical compound

Iron(II) sulfide or ferrous sulfide is one of a family of chemical compounds and minerals with the approximate formula FeS. Iron sulfides are often iron-deficient non-stoichiometric. All are black, water-insoluble solids.

<span class="mw-page-title-main">Sulfur cycle</span> Biogeochemical cycle of sulfur

The important sulfur cycle is a biogeochemical cycle in which the sulfur moves between rocks, waterways and living systems. It is important in geology as it affects many minerals and in life because sulfur is an essential element (CHNOPS), being a constituent of many proteins and cofactors, and sulfur compounds can be used as oxidants or reductants in microbial respiration. The global sulfur cycle involves the transformations of sulfur species through different oxidation states, which play an important role in both geological and biological processes. Steps of the sulfur cycle are:

<span class="mw-page-title-main">Sulfur-reducing bacteria</span> Microorganisms able to reduce elemental sulfur to hydrogen sulfide

Sulfur-reducing bacteria are microorganisms able to reduce elemental sulfur (S0) to hydrogen sulfide (H2S). These microbes use inorganic sulfur compounds as electron acceptors to sustain several activities such as respiration, conserving energy and growth, in absence of oxygen. The final product of these processes, sulfide, has a considerable influence on the chemistry of the environment and, in addition, is used as electron donor for a large variety of microbial metabolisms. Several types of bacteria and many non-methanogenic archaea can reduce sulfur. Microbial sulfur reduction was already shown in early studies, which highlighted the first proof of S0 reduction in a vibrioid bacterium from mud, with sulfur as electron acceptor and H
2
as electron donor. The first pure cultured species of sulfur-reducing bacteria, Desulfuromonas acetoxidans, was discovered in 1976 and described by Pfennig Norbert and Biebel Hanno as an anaerobic sulfur-reducing and acetate-oxidizing bacterium, not able to reduce sulfate. Only few taxa are true sulfur-reducing bacteria, using sulfur reduction as the only or main catabolic reaction. Normally, they couple this reaction with the oxidation of acetate, succinate or other organic compounds. In general, sulfate-reducing bacteria are able to use both sulfate and elemental sulfur as electron acceptors. Thanks to its abundancy and thermodynamic stability, sulfate is the most studied electron acceptor for anaerobic respiration that involves sulfur compounds. Elemental sulfur, however, is very abundant and important, especially in deep-sea hydrothermal vents, hot springs and other extreme environments, making its isolation more difficult. Some bacteria – such as Proteus, Campylobacter, Pseudomonas and Salmonella – have the ability to reduce sulfur, but can also use oxygen and other terminal electron acceptors.

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

Sodium sulfide is a chemical compound with the formula Na2S, or more commonly its hydrate Na2S·9H2O. Both the anhydrous and the hydrated salts in pure crystalline form are colorless solids, although technical grades of sodium sulfide are generally yellow to brick red owing to the presence of polysulfides and commonly supplied as a crystalline mass, in flake form, or as a fused solid. They are water-soluble, giving strongly alkaline solutions. When exposed to moist air, Na2S and its hydrates emit hydrogen sulfide, an extremely toxic, flammable and corrosive gas which smells like rotten eggs.

Anoxic waters are areas of sea water, fresh water, or groundwater that are depleted of dissolved oxygen. The US Geological Survey defines anoxic groundwater as those with dissolved oxygen concentration of less than 0.5 milligrams per litre. Anoxic waters can be contrasted with hypoxic waters, which are low in dissolved oxygen. This condition is generally found in areas that have restricted water exchange.

Microbial metabolism is the means by which a microbe obtains the energy and nutrients it needs to live and reproduce. Microbes use many different types of metabolic strategies and species can often be differentiated from each other based on metabolic characteristics. The specific metabolic properties of a microbe are the major factors in determining that microbe's ecological niche, and often allow for that microbe to be useful in industrial processes or responsible for biogeochemical cycles.

Biogenic sulfide corrosion is a bacterially mediated process of forming hydrogen sulfide gas and the subsequent conversion to sulfuric acid that attacks concrete and steel within wastewater environments. The hydrogen sulfide gas is biochemically oxidized in the presence of moisture to form sulfuric acid. The effect of sulfuric acid on concrete and steel surfaces exposed to severe wastewater environments can be devastating. In the USA alone, corrosion is causing sewer asset losses estimated at $14 billion per year. This cost is expected to increase as the aging infrastructure continues to fail.

Sulfur is metabolized by all organisms, from bacteria and archaea to plants and animals. Sulfur can have an oxidation state from -2 to +6 and is reduced or oxidized by a diverse range of organisms. The element is present in proteins, sulfate esters of polysaccharides, steroids, phenols, and sulfur-containing coenzymes.

<span class="mw-page-title-main">Manure management</span> Agricultural practice

Manure management refers to capture, storage, treatment, and utilization of animal manures in an environmentally sustainable manner. It can be retained in various holding facilities. Animal manure can occur in a liquid, slurry, or solid form. It is utilized by distribution on fields in amounts that enrich soils without causing water pollution or unacceptably high levels of nutrient enrichment. Manure management is a component of nutrient management.

Hydrogen chalcogenides are binary compounds of hydrogen with chalcogen atoms. Water, the first chemical compound in this series, contains one oxygen atom and two hydrogen atoms, and is the most common compound on the Earth's surface.

Euxinia or euxinic conditions occur when water is both anoxic and sulfidic. This means that there is no oxygen (O2) and a raised level of free hydrogen sulfide (H2S). Euxinic bodies of water are frequently strongly stratified; have an oxic, highly productive, thin surface layer; and have anoxic, sulfidic bottom water. The word "euxinia" is derived from the Greek name for the Black Sea (Εὔξεινος Πόντος (Euxeinos Pontos)) which translates to "hospitable sea". Euxinic deep water is a key component of the Canfield ocean, a model of oceans during part of the Proterozoic period (a part specifically known as the Boring Billion) proposed by Donald Canfield, an American geologist, in 1998. There is still debate within the scientific community on both the duration and frequency of euxinic conditions in the ancient oceans. Euxinia is relatively rare in modern bodies of water, but does still happen in places like the Black Sea and certain fjords.

<span class="mw-page-title-main">Microbial oxidation of sulfur</span>

Microbial oxidation of sulfur is the oxidation of sulfur by microorganisms to build their structural components. The oxidation of inorganic compounds is the strategy primarily used by chemolithotrophic microorganisms to obtain energy to survive, grow and reproduce. Some inorganic forms of reduced sulfur, mainly sulfide (H2S/HS) and elemental sulfur (S0), can be oxidized by chemolithotrophic sulfur-oxidizing prokaryotes, usually coupled to the reduction of oxygen (O2) or nitrate (NO3). Anaerobic sulfur oxidizers include photolithoautotrophs that obtain their energy from sunlight, hydrogen from sulfide, and carbon from carbon dioxide (CO2).

<span class="mw-page-title-main">Hydrothermal vent microbial communities</span> Undersea unicellular organisms

The hydrothermal vent microbial community includes all unicellular organisms that live and reproduce in a chemically distinct area around hydrothermal vents. These include organisms in the microbial mat, free floating cells, or bacteria in an endosymbiotic relationship with animals. Chemolithoautotrophic bacteria derive nutrients and energy from the geological activity at Hydrothermal vents to fix carbon into organic forms. Viruses are also a part of the hydrothermal vent microbial community and their influence on the microbial ecology in these ecosystems is a burgeoning field of research.

References

  1. "Hydrogen Sulfide - PubChem Public Chemical Database". The PubChem Project. USA: National Center for Biotechnology Information.
  2. 1 2 Patnaik, Pradyot (2002). Handbook of Inorganic Chemicals. McGraw-Hill. ISBN   978-0-07-049439-8.
  3. 1 2 William M. Haynes (2016). CRC Handbook of Chemistry and Physics (97th ed.). Boca Raton: CRC Press. pp. 4–87. ISBN   978-1-4987-5429-3.
  4. "Hydrogen sulfide". pubchem.ncbi.nlm.nih.gov.
  5. Perrin, D.D. (1982). Ionisation Constants of Inorganic Acids and Bases in Aqueous Solution (2nd ed.). Oxford: Pergamon Press.
  6. 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.
  7. 1 2 Zumdahl, Steven S. (2009). Chemical Principles (6th ed.). Houghton Mifflin Company. p. A23. ISBN   978-0-618-94690-7.
  8. "Hydrogen sulfide". npi.gov.au.
  9. 1 2 "Hydrogen sulfide". Immediately Dangerous to Life or Health Concentrations (IDLH). National Institute for Occupational Safety and Health (NIOSH).
  10. 1 2 3 NIOSH Pocket Guide to Chemical Hazards. "#0337". National Institute for Occupational Safety and Health (NIOSH).
  11. Greenwood, Norman N.; Earnshaw, Alan (1997). Chemistry of the Elements (2nd ed.). Butterworth-Heinemann. ISBN   978-0-08-037941-8.
  12. Smith, Roger P. (2010). "A Short History of Hydrogen Sulfide". American Scientist. 98 (1): 6. doi:10.1511/2010.82.6.
  13. Shackelford, R. E.; Li, Y.; Ghali, G. E.; Kevil, C. G. (2021). "Bad Smells and Broken DNA: A Tale of Sulfur-Nucleic Acid Cooperation". Antioxidants. 10 (11): 1820. doi: 10.3390/antiox10111820 . PMC   8614844 . PMID   34829691.
  14. Reiffenstein, R. J.; Hulbert, W. C.; Roth, S. H. (1992). "Toxicology of Hydrogen Sulfide". Annual Review of Pharmacology and Toxicology. 32: 109–134. doi:10.1146/annurev.pa.32.040192.000545. PMID   1605565.
  15. 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.
  16. 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.
  17. Drozdov, A.; Eremets, M. I.; Troyan, I. A. (2014). "Conventional superconductivity at 190 K at high pressures". arXiv: 1412.0460 [cond-mat.supr-con].
  18. Cartlidge, Edwin (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.
  19. Faraji, F. (1998). "The direct conversion of hydrogen sulfide to hydrogen and sulfur". International Journal of Hydrogen Energy. 23 (6): 451–456. doi:10.1016/S0360-3199(97)00099-2.
  20. JCE staff (March 2000). "Silver to Black - and Back". Journal of Chemical Education. 77 (3): 328A. Bibcode:2000JChEd..77R.328J. doi:10.1021/ed077p328a. ISSN   0021-9584.
  21. "What causes coins to tone – ICCS". iccscoin.ca. Retrieved 2024-02-11.
  22. "Coin Toning 101: The Differences between Naturally and Artificially Toned Coins". Original Skin Coins. Retrieved 2021-10-15.
  23. 1 2 3 4 Pouliquen, Francois; Blanc, Claude; Arretz, Emmanuel; Labat, Ives; Tournier-Lasserve, Jacques; Ladousse, Alain; Nougayrede, Jean; Savin, Gérard; Ivaldi, Raoul; Nicolas, Monique; Fialaire, Jean; Millischer, René; Azema, Charles; Espagno, Lucien; Hemmer, Henri; Perrot, Jacques (2000). "Hydrogen Sulfide". Ullmann's Encyclopedia of Industrial Chemistry. doi:10.1002/14356007.a13_467. ISBN   3527306730.
  24. Chan, Yi Herng; Loy, Adrian Chun Minh; Cheah, Kin Wai; Chai, Slyvester Yew Wang; Ngu, Lock Hei; How, Bing Shen; Li, Claudia; Lock, Serene Sow Mun; Wong, Mee Kee; Yiin, Chung Loong; Chin, Bridgid Lai Fui; Chan, Zhe Phak; Lam, Su Shiung (2023). "Hydrogen sulfide (H2S) conversion to hydrogen (H2) and value-added chemicals: Progress, challenges and outlook" (PDF). Chemical Engineering Journal. 458. doi:10.1016/j.cej.2023.141398. S2CID   255887336.
  25. McPherson, William (1913). Laboratory manual. Boston: Ginn and Company. p. 445.
  26. 1 2 Huang, Caleb Weihao; Moore, Philip Keith (2015), "H2S Synthesizing Enzymes: Biochemistry and Molecular Aspects", Chemistry, Biochemistry and Pharmacology of Hydrogen Sulfide, Handbook of Experimental Pharmacology, Springer International Publishing, vol. 230, pp. 3–25, doi:10.1007/978-3-319-18144-8_1, ISBN   9783319181431, PMID   26162827
  27. 1 2 3 Kabil, Omer; Vitvitsky, Victor; Xie, Peter; Banerjee, Ruma (15 July 2011). "The Quantitative Significance of the Transsulfuration Enzymes for H 2 S Production in Murine Tissues". Antioxidants & Redox Signaling. 15 (2): 363–372. doi:10.1089/ars.2010.3781. PMC   3118817 . PMID   21254839.
  28. 1 2 3 Kabil, Omer; Banerjee, Ruma (10 February 2014). "Enzymology of H2S Biogenesis, Decay and Signaling". Antioxidants & Redox Signaling. 20 (5): 770–782. doi:10.1089/ars.2013.5339. PMC   3910450 . PMID   23600844.
  29. "Why Does My Water Smell Like Rotten Eggs? Hydrogen Sulfide and Sulfur Bacteria in Well Water". Minnesota Department of Health. Archived from the original on 11 March 2015. Retrieved 1 December 2014.
  30. Wallace, John L.; Wang, Rui (May 2015). "Hydrogen sulfide-based therapeutics: exploiting a unique but ubiquitous gasotransmitter". Nature Reviews Drug Discovery. 14 (5): 329–345. doi:10.1038/nrd4433. PMID   25849904. S2CID   5361233.
  31. Powell, Chadwick R.; Dillon, Kearsley M.; Matson, John B. (2018). "A review of hydrogen sulfide (H2S) donors: Chemistry and potential therapeutic applications". Biochemical Pharmacology. 149: 110–123. doi:10.1016/j.bcp.2017.11.014. ISSN   0006-2952. PMC   5866188 . PMID   29175421.
  32. 1 2 3 4 5 Hancock, John T. (2017). Cell signalling (Fourth ed.). Oxford, United Kingdom. ISBN   9780199658480. OCLC   947925636.{{cite book}}: CS1 maint: location missing publisher (link)
  33. 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. PMC   5862769 . PMID   28947277.
  34. 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. PMC   5629631 . PMID   28978633.
  35. 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. PMID   22535897. S2CID   21932297.
  36. 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 . PMID   30355137. S2CID   53027283.
  37. Kimura, Hideo (February 2020). "Signalling by Hydrogen Sulfide and Polysulfides via Protein S-Sulfuration". British Journal of Pharmacology. 177 (4): 720–733. doi:10.1111/bph.14579. PMC   7024735 . PMID   30657595.
  38. Lee, J. D. (1998). Concise inorganic chemistry (5. ed., reprinted ed.). Oxford: Blackwell Science. p. 538. ISBN   978-0-632-05293-6.
  39. Khazaei, Ardeshir; Kazem-Rostami, Masoud; Moosavi-Zare, Ahmad; Bayat, Mohammad; Saednia, Shahnaz (August 2012). "Novel One-Pot Synthesis of Thiophenols from Related Triazenes under Mild Conditions". Synlett. 23 (13): 1893–1896. doi:10.1055/s-0032-1316557. S2CID   196805424.
  40. 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.
  41. "Hydrogen Sulphide In Well Water" . Retrieved 4 September 2018.
  42. Agency for Toxic Substances and Disease Registry (July 2006). "Toxicological Profile For Hydrogen Sulfide" (PDF). p. 154. Retrieved 2012-06-20.
  43. OnePetro. "Home - OnePetro". onepetro.org. Archived from the original on 2013-10-14. Retrieved 2013-08-14.
  44. "Hydrogen Sulfide" (PDF). Agency for Toxic Substances and Disease Registry. December 2016.
  45. 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.
  46. Cavinato, C. (2013) [2013]. "Anaerobic digestion fundamentals" (PDF).
  47. Pokorna, Dana; Zabranska, Jana (November 2015). "Sulfur-oxidizing bacteria in environmental technology". Biotechnology Advances. 33 (6): 1246–1259. doi:10.1016/j.biotechadv.2015.02.007. PMID   25701621.
  48. Lemley, Ann T.; Schwartz, John J.; Wagenet, Linda P. "Hydrogen Sulfide in Household Drinking Water" (PDF). Cornell University. Archived from the original (PDF) on 19 August 2019.
  49. "Hydrogen Sulfide (Rotten Egg Odor) in Pennsylvania Groundwater Wells". Penn State. Penn State College of Agricultural Sciences. Archived from the original on 4 January 2015. Retrieved 1 December 2014.
  50. McFarland, Mark L.; Provin, T. L. "Hydrogen Sulfide in Drinking Water Treatment Causes and Alternatives" (PDF). Texas A&M University. Retrieved 1 December 2014.
  51. 1 2 3 4 5 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. Archived from the original on June 15, 2013. Retrieved 2013-06-07.{{cite journal}}: CS1 maint: unfit URL (link)
  52. Harrison, J. Bower (1843-11-18). "Some Remarks on the Production of Sulphuretted Hydrogen Gas in the Alimentary Canal, and Its Effects on the System". Provincial Medical Journal and Retrospect of the Medical Sciences, BMJ. 7 (164): 127–129. JSTOR   25492480.
  53. "A Short History of Hydrogen Sulfide". American Scientist. 2017-02-06. Retrieved 2023-12-25.
  54. "Why Does My Water Smell Like Rotten Eggs?". Minnesota Department of Health. Retrieved 20 January 2020.
  55. Contaminants, National Research Council (US) Committee on Emergency and Continuous Exposure Guidance Levels for Selected Submarine (2009). Hydrogen Sulfide. National Academies Press (US).
  56. Iowa State University. "Hydrogen Sulfide Material Safety Data Sheet" (PDF). Department of Chemistry. Archived from the original (PDF) on 2009-03-27. Retrieved 2009-03-14.
  57. Ramasamy, S.; Singh, S.; Taniere, P.; Langman, M. J. S.; Eggo, M. C. (August 2006). "Sulfide-detoxifying enzymes in the human colon are decreased in cancer and upregulated in differentiation". American Journal of Physiology. Gastrointestinal and Liver Physiology. 291 (2): G288–G296. doi:10.1152/ajpgi.00324.2005. PMID   16500920. S2CID   15443357.
  58. Lewis, R.J. (1996). Sax's Dangerous Properties of Industrial Materials (9th ed.). New York, NY: Van Nostrand Reinhold.[ page needed ]
  59. 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. Bibcode:1982IAOEH..51...55H. doi:10.1007/bf00378410. PMID   7152702. S2CID   2768183.
  60. "The chemical suicide phenomenon". Firerescue1.com. 2011-02-07. Retrieved 2013-12-19.
  61. "Fatal chemical inhalations in the workplace up in 2017". U.S. Bureau of Labor Statistics. Retrieved 15 April 2022.
  62. "Hydrogen Sulfide - Hazards | Occupational Safety and Health Administration". www.osha.gov. Retrieved 2021-09-27.
  63. Iowa State University Extension (May 2004). "The Science of Smell Part 1: Odor perception and physiological response" (PDF). PM 1963a. Retrieved 2012-06-20.
  64. USEPA; Health and Environmental Effects Profile for Hydrogen Sulfide p.118-8 (1980) ECAO-CIN-026A
  65. Zenz, C.; Dickerson, O.B.; Horvath, E.P. (1994). Occupational Medicine (3rd ed.). St. Louis, MO. p. 886.{{cite book}}: CS1 maint: location missing publisher (link)
  66. Gerasimon, Gregg; Bennett, Steven; Musser, Jeffrey; Rinard, John (January 2007). "Acute hydrogen sulfide poisoning in a dairy farmer". Clinical Toxicology. 45 (4): 420–423. doi:10.1080/15563650601118010. PMID   17486486. S2CID   10952243.
  67. 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.
  68. Hsu, P; Li, H-W; Lin, Y-T (1987). "Acute hydrogen sulfide poisoning treated with hyperbaric oxygen". Journal of Hyperbaric Medicine. 2 (4): 215–221. Archived from the original on December 7, 2008.{{cite journal}}: CS1 maint: unfit URL (link)
  69. 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.
  70. "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.
  71. Becerra, Hector; Pierson, David (2005-09-03). "Gas Kills 3 Crewmen on Ship". Los Angeles Times.
  72. Ferguson, Dan (September 16, 2011). "Details of Langley mushroom farm tragedy finally disclosed". Abbotsford News . Retrieved April 13, 2020.
  73. 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.
  74. "Do not breathe: Dangerous, toxic gas found at Siam Square One". Coconuts Bangkok. Coconuts Media. 2014-10-21. Retrieved 20 November 2014.
  75. "Two sewer workers die, apparently due to toxic fumes - CBS News". CBS News . 26 August 2014.
  76. "Russian capital Moscow shrouded in noxious gas". BBC News. British Broadcasting Corporation. 2014-11-10. Retrieved 1 December 2014.
  77. "Sources: Mom, daughter found dead in Porsche likely died from carbon monoxide". WFTV. 7 June 2016. Both had red skin and rash-like symptoms, and had vomited, sources said.
  78. Salinger, Tobias (4 October 2016). "Woman, girl died after inhaling hydrogen sulfide, coroners say". New York Daily News . Retrieved 28 April 2017.
  79. 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.
  80. Kealing, Bob. "Medical examiner confirms suspected cause of deaths in Turnpike mystery". Archived from the original on 2016-10-05. Retrieved 2016-10-04.
  81. 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.
  82. "One by one, 3 utility workers descended into a manhole. One by one, they died". www.washingtonpost.com. Archived from the original on 2017-01-18.
  83. 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.
  84. "Key Largo firefighter takes first steps since nearly getting killed". WSVN. 18 January 2017.
  85. "Firefighter who survived Key Largo rescue attempt that killed 3 leaves hospital". Sun Sentinel. The Associated Press. 26 January 2017.
  86. Rabin, Charles; Goodhue, David (16 January 2017). "Three Keys utility workers die in wastewater trench". Miami Herald . Retrieved 28 April 2017.
  87. 1 2 Clantar, Claire (25 Sep 2020). "Former Victorian paper mill fined $1 million after deaths of two workers". 9News. Retrieved 30 May 2021.
  88. "Two fatalities in suspected hydrogen sulfide gas leak at paper mill". Australian Institute of Health & Safety. 31 May 2018. Retrieved 30 May 2021.
  89. Brescia, Paul (28 May 2018). "SafeWork investigating Norske Skog". Sprinter. Retrieved 30 May 2021.
  90. SafeWork NSW v Norske Skog Paper Mills (Australia) Limited, NSWDC 559 ( District Court of New South Wales 25 September 2020).
  91. SafeWork NSW (29 Mar 2021). Incident Animation – Hazardous Gas (Motion picture). Archived from the original on 2021-10-30. Retrieved 30 May 2021.
  92. "Feds Probe Fatal 2019 Hydrogen Sulfide Release in Texas". Industrial Fire World. 27 July 2020. Retrieved 29 May 2021.
  93. "Aghorn Operating Waterflood Station Hydrogen Sulfide Release". U.S. Chemical Safety and Hazardous Investigation Board. 21 May 2021. Retrieved 29 May 2021.
  94. "Dangerous Japanese 'Detergent Suicide' Technique Creeps Into U.S". Wired. March 13, 2009.
  95. Namiki, Noriko (23 May 2008). "Terrible Twist in Japan Suicide Spates". ABC News.
  96. http://info.publicintelligence.net/LARTTAChydrogensulfide.pdf%5B%5D
  97. http://info.publicintelligence.net/MAchemicalsuicide.pdf%5B%5D
  98. http://info.publicintelligence.net/illinoisH2Ssuicide.pdf%5B%5D
  99. http://info.publicintelligence.net/NYhydrogensulfide.pdf%5B%5D
  100. http://info.publicintelligence.net/KCTEWhydrogensulfide.pdf%5B%5D
  101. "Chemical Suicide on Campus" (PDF). www.maryland.gov. Archived from the original on January 3, 2012.
  102. Scoville, Dean (April 2011). "Chemical Suicides". POLICE Magazine. Retrieved 2013-12-19.
  103. Casey, Connor (26 May 2020). "Hana Kimura Cause Of Death Revealed". ComicBook.com. More details about her death have since come to light, as Dave Meltzer provided details about what happened on the night of her death during a recent Wrestling Observer Radio. According to Meltzer, Kimura died after inhalation of hydrogen sulfide. He explained that concerns about her health first popped up when she posted a tweet early Saturday morning indicating that she was going to cause self-harm.
  104. Barton, Larry L.; Fardeau, Marie-Laure; Fauque, Guy D. (2014). "Hydrogen Sulfide: A Toxic Gas Produced by Dissimilatory Sulfate and Sulfur Reduction and Consumed by Microbial Oxidation". The Metal-Driven Biogeochemistry of Gaseous Compounds in the Environment. Metal Ions in Life Sciences. Vol. 14. pp. 237–277. doi:10.1007/978-94-017-9269-1_10. ISBN   978-94-017-9268-4. PMID   25416397.
  105. 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.
  106. Wächtershäuser, G (December 1988). "Before enzymes and templates: theory of surface metabolism". Microbiological Reviews. 52 (4): 452–484. doi:10.1128/MMBR.52.4.452-484.1988. PMC   373159 . PMID   3070320.
  107. 1 2 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. Bibcode:2008JFBio..72..523T. doi:10.1111/j.1095-8649.2007.01716.x. S2CID   27303725.
  108. 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.
  109. "Hydrothermal Vents". Marine Society of Australia. Retrieved 28 December 2014.
  110. 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.
  111. Kumaresan, Deepak; Wischer, Daniela; Stephenson, Jason; Hillebrand-Voiculescu, Alexandra; Murrell, J. Colin (16 March 2014). "Microbiology of Movile Cave—A Chemolithoautotrophic Ecosystem". Geomicrobiology Journal. 31 (3): 186–193. Bibcode:2014GmbJ...31..186K. doi:10.1080/01490451.2013.839764. S2CID   84472119.
  112. Despois, D. (1997). "Radio Line Observations Of Molecular And Isotopic Species In Comet C/1995 O1 (Hale-Bopp)". Earth, Moon, and Planets. 79 (1/3): 103–124. Bibcode:1997EM&P...79..103D. doi:10.1023/A:1006229131864. S2CID   118540103.
  113. Irwin, Patrick G. J.; Toledo, Daniel; Garland, Ryan; Teanby, Nicholas A.; Fletcher, Leigh N.; Orton, Glenn A.; Bézard, Bruno (May 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 . S2CID   102775371.
  114. 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.[ page needed ]
  115. 1 2 "Impact from the Deep". Scientific American. October 2006.
  116. Lamarque, J.-F.; Kiehl, J. T.; Orlando, J. J. (16 January 2007). "Role of hydrogen sulfide in a Permian-Triassic boundary ozone collapse". Geophysical Research Letters . 34 (2): 1–4. Bibcode:2007GeoRL..34.2801L. doi: 10.1029/2006GL028384 . S2CID   55812439.
  117. Kump, Lee; Pavlov, Alexander; Arthur, Michael A. (1 May 2005). "Massive release of hydrogen sulfide to the surface ocean and atmosphere during intervals of oceanic anoxia". Geology . 33 (5): 397–400. Bibcode:2005Geo....33..397K. doi:10.1130/G21295.1 . Retrieved 2 April 2023.

Additional resources