Fact-checked by Grok 2 weeks ago

Hydroxy group

The hydroxy group, also known as the hydroxyl group and denoted as -OH, is a fundamental in consisting of an oxygen atom covalently bonded to a and attached to a carbon atom within an . It serves as the defining characteristic of alcohols, where it is represented by the general formula ROH, with R denoting an alkyl or . This group imparts distinct chemical and physical properties to the molecules it is part of, distinguishing them from hydrocarbons of similar size. The hydroxy group's polarity arises from the significant difference between oxygen and , making the O-H highly polar and capable of forming bonds as both a donor (via the ) and an acceptor (via the oxygen's lone pairs). These hydrogen-bonding capabilities lead to elevated boiling and melting points for hydroxy-containing compounds compared to nonpolar hydrocarbons of equivalent molecular weight, as well as enhanced in . In terms of reactivity, the hydroxy group can undergo to form alkoxides under basic conditions, participate in oxidation reactions to yield carbonyl compounds, and serve as a site for esterification or formation, influencing the and behavior of materials. Hydroxy groups are prevalent in biological systems, appearing in carbohydrates (such as glucose with multiple -OH groups enabling glycosidic linkages), amino acids like serine and threonine (where the side-chain hydroxy facilitates phosphorylation), and lipids including steroids. Their hydrogen-bonding prowess is vital for maintaining protein secondary structures, stabilizing DNA base pairs indirectly through water interactions, and enabling enzymatic reactions. In medicinal chemistry, hydroxy groups contribute to drug efficacy by enhancing solubility, binding affinity to targets, and metabolic stability; notably, they are present in approximately 40% of approved drugs and 65% of known natural products. Nomenclature for compounds bearing this group prioritizes it highly, using the suffix -ol for alcohols (e.g., ethanol) or the prefix hydroxy- when another higher-priority functional group is present.

Definition and Fundamentals

Chemical Definition

The hydroxy group, denoted as −OH, is a composed of an oxygen atom covalently bonded to a . In , it is typically attached to a carbon atom, forming alcohols with the general structure R−OH, where R represents an alkyl or . This attachment imparts specific reactivity to the due to the polar of the O−H bond. In , the hydroxy group can bond to metals or nonmetals, as in compounds like aluminum hydroxide (Al(OH)3) or (HOCl), where it functions as a or part of an oxoacid structure. The hydroxy group differs fundamentally from related species such as the hydroxide ion (OH) and the hydroxyl radical (OH•). The hydroxide ion is an anionic species in which the oxygen atom is bonded to hydrogen and carries a formal negative charge, with eight valence electrons around oxygen including three lone pairs; it typically exists in ionic compounds or aqueous solutions. In contrast, the hydroxyl radical is a neutral, highly reactive species featuring an unpaired electron on the oxygen atom, resulting in seven valence electrons around oxygen and only two lone pairs, making it a free radical rather than a stable functional group. The hydroxy group itself involves covalent bonding where oxygen shares electrons with both the attached atom (e.g., carbon) and hydrogen, achieving a stable octet without charge or unpaired electrons. The recognition of the hydroxy group in organic compounds traces back to the late , when conducted combustion analyses that revealed the elemental composition of alcohols like (spirit of wine), identifying it as containing , , and in a specific ratio consistent with the R−OH structure. These experiments laid the groundwork for understanding functional groups in by shifting focus from qualitative observations to quantitative .

Nomenclature

In organic chemistry, the hydroxy group (-OH) in alcohols is named using substitutive nomenclature according to IUPAC recommendations, where the parent hydride chain is selected and modified by replacing the ending "e" with the suffix "-ol" to indicate the principal characteristic group. The position of the hydroxy group is specified by a locant, with the chain numbered to give the lowest possible number to the carbon atom bearing the -OH. For example, CH₃CH₂OH is named ethanol, reflecting its two-carbon parent chain with the hydroxy group at position 1. In cases of multiple hydroxy groups, the suffix becomes "-diol," "-triol," etc., with locants for each, such as 1,2-ethanediol for HOCH₂CH₂OH. When the hydroxy group is not the principal function, it is expressed as the prefix "hydroxy-." In inorganic chemistry, nomenclature distinguishes between ionic compounds containing hydroxide ions (OH) and coordination compounds with covalent hydroxy ligands. Simple ionic metal hydroxides, such as NaOH, are named by combining the cation name with "hydroxide," yielding sodium hydroxide. For coordination compounds, the covalent hydroxy ligand is named "hydroxido," listed alphabetically before the central atom, which includes its oxidation state in parentheses; for instance, [Fe(OH)(H2O)5]²⁺ is pentaaquahydroxidoiron(3+). This distinguishes ionic hydroxides from covalent hydroxy ligands in complexes. Special cases arise for phenols and enols. Phenols, where the hydroxy group is directly attached to an aromatic ring (Ar-OH), are named as derivatives of the retained name "phenol," such as 2-bromophenol for the ortho-substituted compound, with the parent structure being numbered from the carbon attached to -OH. Enols, featuring a hydroxy group on a carbon-carbon and often existing as tautomers of carbonyl compounds, are named substitutively with the "-ol" suffix and an "en" for the , like prop-1-en-2-ol for the form of acetone. The of hydroxy compounds has undergone significant evolution, particularly in the , transitioning from descriptive systems to standardized approaches under IUPAC. Early names like "carbinol" for derivatives (e.g., dimethyl carbinol for propan-2-ol) were common in the 19th and early 20th centuries but were largely replaced by the systematic "-ol" suffix following the establishment of IUPAC rules in the and refinements through the . This shift, formalized in documents like the 1957 and 1971 IUPAC recommendations, emphasized unambiguous, generative naming based on parent structures and functional priorities.

Physical and Chemical Properties

Bonding and Structure

The hydroxy group (-) consists of an oxygen atom to a and typically to a carbon atom in compounds such as alcohols. The O-H is approximately 0.96 , as determined from experimental data on simple alcohols like and . In alcohols, the C-O is about 1.43 , reflecting a single between the carbon and oxygen atoms. The bond angle around the carbon atom attached to the -OH group approaches 109°, consistent with tetrahedral geometry in sp³-hybridized carbon centers. The oxygen atom in the hydroxy group of alcohols is sp³ hybridized, forming four equivalent hybrid orbitals: two used for sigma bonds to carbon and , and the remaining two occupied by lone pairs of electrons. This hybridization results in a bent structure at the oxygen, with the C-O-H bond angle slightly less than 109° due to repulsion from the lone pairs. In , where the -OH group is directly attached to an aromatic ring, one of the oxygen's lone pairs participates in delocalization into the pi system of the benzene ring, leading to partial double-bond character in the C-O bond and altered bond orders throughout the ring. Spectroscopically, the hydroxy group exhibits characteristic signatures that reveal its bonding. In infrared (IR) spectroscopy, the O-H stretching appears as a broad absorption band between 3200 and 3600 cm⁻¹, arising from the polar O-H bond and influenced by molecular environment. In nuclear magnetic resonance (¹H NMR) spectroscopy, the -OH proton signal is variable, typically ranging from 1 to 5 ppm, due to rapid proton exchange that averages the and often broadens the peak.

Polarity and Intermolecular Interactions

The hydroxy group (-OH) exhibits significant polarity due to the electronegativity difference between oxygen (3.44) and hydrogen (2.20) on the Pauling scale, resulting in a partial negative charge (δ⁻) on the oxygen atom and a partial positive charge (δ⁺) on the hydrogen atom. This charge separation creates a polar covalent O-H bond, with the dipole moment enhancing the group's ability to participate in intermolecular interactions. The oxygen atom's two lone pairs of electrons, arising from its bonding structure, further facilitate these interactions by allowing the hydroxy group to act both as a hydrogen bond donor (via the δ⁺ hydrogen) and acceptor (via the δ⁻ oxygen). Hydrogen bonding is the primary intermolecular interaction involving the hydroxy group, where the partially positive of one - forms an attractive force with the partially negative oxygen of another - or a similar group. These bonds typically have energies around 20 kJ/mol, as seen in dimers (approximately 20.8 kJ/mol) and alcohol systems like . In , each can form up to four hydrogen bonds, creating an extensive network that influences molecular arrangement. Similarly, in alcohols such as and , the - groups enable both intra- and intermolecular hydrogen bonding, leading to associated structures in the liquid state. This polarity and hydrogen bonding significantly elevate the boiling points of compounds containing hydroxy groups compared to non-polar analogs of similar molecular weight. For instance, (C₂H₅OH) has a of 78°C, much higher than (C₃H₈) at -42°C, due to the required to disrupt hydrogen bonds during vaporization. Hydrogen bonding also enhances in ; small alcohols like and are fully miscible, as their -OH groups form hydrogen bonds with molecules, overcoming hydrophobic effects from alkyl chains. In comparison to other functional groups, the hydroxy group's hydrogen bonding is stronger than that in amines or ethers. Amines form weaker N-H···N bonds (due to nitrogen's lower of 3.04), resulting in lower boiling points for similar-sized compounds, such as (16°C) versus (78°C)./03:_An_Introduction_to_Organic_Compounds_Structure_and_Nomenclature/3.07:_The_Physical_Properties_of_Alkanes_Alkyl_Halides_Alcohols_Ethers_and_Amines) Ethers, lacking a hydrogen donor, rely only on dipole-dipole interactions and exhibit even lower boiling points, like (35°C), underscoring the hydroxy group's superior intermolecular attraction./03:_An_Introduction_to_Organic_Compounds_Structure_and_Nomenclature/3.07:_The_Physical_Properties_of_Alkanes_Alkyl_Halides_Alcohols_Ethers_and_Amines)

Acidity and Reactivity

The hydroxy group in alcohols exhibits weak acidity, with pK_a values typically ranging from 15 to 18, making challenging under standard conditions. For example, has a pK_a of 15.9, and has a pK_a of 15.5. In contrast, the hydroxy group attached to an aromatic ring, as in , is more acidic, with a pK_a around 10, due to stabilization of the resulting phenoxide , where the negative charge is delocalized into the ring. This delocalization enhances the stability of the conjugate base compared to the ions from aliphatic alcohols. The basicity of the hydroxy group is quite weak, as protonation occurs on the oxygen atom to form the oxonium ion R-OH₂⁺, which has a pK_a of approximately -2, indicating that such species are strong acids and poor bases in neutral environments. This low basicity arises from the high electronegativity of oxygen, which holds electron density tightly, limiting its ability to accept protons effectively. Substituents adjacent to the hydroxy group can significantly influence its acidity through electronic effects. Electron-withdrawing groups stabilize the conjugate base by dispersing the negative charge, thereby lowering the pK_a; for instance, has a pK_a of 12.4 due to the inductive withdrawal by the trifluoromethyl group. Electron-donating substituents, conversely, increase the pK_a by destabilizing the anion. In terms of reactivity, the oxygen atom in the hydroxy group serves as a in bimolecular (S_N2) reactions, where it attacks centers such as alkyl halides, though its nucleophilicity is moderated by and the need for in some cases. The attached to oxygen, meanwhile, acts as an in processes, facilitating acid-base reactions with stronger bases. Hydrogen bonding in protic solvents can modestly enhance the effective acidity of the hydroxy group by stabilizing the for proton transfer.

Occurrence and Applications

In Organic Compounds

In organic compounds, the hydroxy group (-OH) is a key that imparts distinct chemical properties when attached to carbon atoms, primarily manifesting in classes such as alcohols, polyols, , and enols. Alcohols, characterized by the -OH group bonded to a saturated carbon atom, are foundational examples and are classified based on the substitution at the carbon bearing the hydroxy group. Primary alcohols have the -OH attached to a carbon bonded to only one other carbon atom, such as methanol (CH_3OH) and ethanol (CH_3CH_2OH), which exhibit relatively higher reactivity in oxidation processes. Secondary alcohols feature the -OH on a carbon bonded to two other carbons, exemplified by isopropanol ((CH_3)_2CHOH), while tertiary alcohols have the -OH on a carbon bonded to three other carbons, like tert-butanol ((CH_3)_3COH). This classification influences their physical properties and reactivity patterns, with tertiary alcohols generally showing greater steric hindrance. Polyols are compounds containing multiple hydroxy groups attached to carbon atoms, enhancing their and hydrogen-bonding capabilities. Diols possess two -OH groups, such as (HOCH_2CH_2OH), and triols have three, as seen in (HOCH_2CH(OH)CH_2OH), which serves as a structural component in various lipid-related molecules. These multiple hydroxy groups allow polyols to form extensive intermolecular interactions, contributing to their roles in molecular assembly. Phenols feature the hydroxy group directly attached to an aromatic ring, such as in phenol (C_6H_5OH), where resonance delocalization with the benzene ring increases acidity compared to aliphatic alcohols. In contrast, enols contain the -OH group bonded to a vinylic carbon in a C=C-OH arrangement, often existing in equilibrium with keto forms through tautomerism in carbonyl compounds like acetone enol (CH_2=C(OH)CH_3). This tautomerism stabilizes the enol form in certain systems, particularly those conjugated with aromatic rings. The hydroxy group in alcohols undergoes interconversions, such as oxidation, where primary alcohols are converted to aldehydes and then carboxylic acids, and secondary alcohols to ketones, using oxidizing agents like . Tertiary alcohols resist oxidation under mild conditions due to the absence of a on the hydroxy-bearing carbon. of alcohols, typically under acidic conditions, eliminates to form s, with secondary and tertiary alcohols yielding more stable alkene products via intermediates. The hydroxy group's ability to form bonds significantly enhances the of lower-molecular-weight alcohols and polyols.

In Inorganic and Biochemical Contexts

In inorganic chemistry, the hydroxy group commonly appears in metal hydroxides, where it serves as a ligand bound to metal cations, often forming ionic or layered structures. For instance, calcium hydroxide, Ca(OH)₂, consists of Ca²⁺ ions coordinated to two OH⁻ anions, resulting in a compound with strong basic properties due to the dissociation of the hydroxide ions in aqueous solution, yielding a pH of approximately 12.5–12.8. This structure exemplifies the role of hydroxy groups in stabilizing metal centers through coordinate bonds, contributing to the material's use in applications like pH adjustment, though here the focus is on its fundamental coordination. Similarly, in coordination chemistry, hydroxy groups participate in aquo-hydroxy complexes, such as the tetrahedral aluminate ion [Al(OH)₄]⁻, which predominates in alkaline solutions where aluminum exists as a tetrahedrally coordinated species with four hydroxy ligands. These complexes illustrate how hydroxy groups can bridge or terminal-ligate metal ions, influencing speciation and stability in aqueous environments, as seen in the transition from octahedral aquo species to hydroxy-dominant forms under basic conditions. In biochemical contexts, hydroxy groups are integral to the structure and function of biomolecules beyond carbon-based frameworks. In , the side-chain hydroxy groups of serine and enable key interactions; for example, the -CH₂OH in serine and -CH(OH)CH₃ in facilitate hydrogen bonding and serve as sites for post-translational modifications. These hydroxy moieties are critical in active sites, where they participate in proton transfer mechanisms, such as in serine proteases like , in which the serine hydroxyl acts as a after facilitated by a , enabling of peptide bonds. In carbohydrates, multiple hydroxy groups on glucose—such as those at C-1, C-2, C-3, C-4, and C-6—contribute to the molecule's solubility and reactivity through hydrogen bonding, allowing glucose to form structural like or serve as an energy source in . Additionally, in cofactors like ascorbic acid (), the enediol hydroxy groups at C-2 and C-3 enable its role as a and cofactor, donating electrons in reactions, such as collagen synthesis, due to the proximity of these groups to carbonyl functionalities. The biological functions of hydroxy groups extend to dynamic processes like and signaling. In , glycosyltransferases attach sugar moieties to the hydroxy groups of or residues on proteins, modulating , stability, and interactions in cellular pathways. This modification plays a pivotal role in , as glycosylated proteins on the cell surface influence receptor-ligand interactions and , such as in pathways where altered glycosylation affects receptor and downstream effects like . Regarding pH-dependent speciation, hydroxy groups in these biomolecules, with values typically ranging from 15 to 18 for aliphatic alcohols, remain predominantly protonated (as R-OH) under physiological conditions at 7.4, limiting but allowing participation in networks without . This protonated state ensures stability while enabling reversible proton transfer in enzymatic reactions, contrasting with more acidic hydroxy groups like in (pKa ≈ 10), which can partially deprotonate near neutral .

Industrial and Synthetic Uses

The hydroxy group plays a central role in industrial synthesis of alcohols through methods like the of alkenes and the of carbonyl compounds, leveraging its ability to form stable bonds in reactions. One prominent industrial process is the indirect of to produce , where reacts with concentrated to form ethyl hydrogen sulfate, followed by with to yield and regenerate the acid; this method, though largely superseded by direct using catalysts at high temperatures (around 300°C) and pressures (60-70 atm), remains relevant for understanding early routes. Direct , employed by major producers, achieves yields up to 95% and is economically viable due to the abundance of from cracking. Another key synthetic route involves the of carbonyl groups in aldehydes and ketones to primary and secondary alcohols, respectively; industrially, this is predominantly achieved via catalytic using gas and metal catalysts like or under moderate pressures (10-50 atm) and temperatures (100-200°C), as seen in the production of fatty alcohols from aldehydes derived from natural fats or synthetic sources, offering high selectivity and scalability over lab-scale reagents like NaBH₄. Ethanol, a bearing a , exemplifies the economic significance of these syntheses, serving as a versatile , , and chemical intermediate with global production reaching approximately 110 billion liters (about 87 million metric tons) annually in the early , driven largely by demand in the United States and for blended into . Glycerol, a with three , is another critical industrial product obtained as a byproduct of from vegetable oils or animal fats via , and it finds applications in the manufacture of for explosives—where its react with to form the —and in pharmaceuticals as a , , and stabilizer in formulations like cough syrups and suppositories. In broader applications, hydroxy-containing compounds like fatty alcohols (e.g., lauryl and cetyl alcohols) are essential raw materials for non-ionic surfactants in detergents, shampoos, and industrial cleaners, where the hydroxy group facilitates to produce emulsifiers that reduce and enhance properties, with global demand exceeding 3 million tons per year due to their role in household and . In the pharmaceutical sector, the hydroxy group in compounds like (acetaminophen) contributes to its and activity by enabling hydrogen bonding interactions with biological targets; is synthesized industrially via of p-aminophenol, a process yielding over 100,000 tons annually worldwide for pain relief medications. The reactivity of the hydroxy group in these syntheses and applications stems from its nucleophilicity and ability to participate in hydrogen bonding, which underpins their utility in forming esters, ethers, and other derivatives. From an environmental perspective, aliphatic alcohols such as and exhibit high biodegradability, with over 90% degradation in aerobic within days via microbial oxidation of the hydroxy group to carboxylic acids, facilitating their use in eco-friendly formulations; in contrast, hydroxy compounds like those in substituted can persist longer in the due to their aromatic stability and potential formation of toxic intermediates during , necessitating advanced treatment in industrial effluents to mitigate ecotoxicity.

Hydroxyl Radical

Structure and Formation

The (OH•) is a diatomic free radical composed of a covalently bonded to an oxygen atom, featuring an primarily localized on the oxygen, resulting in a ground electronic state of ^2\Pi. This structure imparts high reactivity, distinguishing it from the stable, closed-shell (-OH) found in molecules like alcohols. The O-H in the is 0.97 Å, while the , which corresponds to the energy required to break the bond into ground-state H and O atoms, is 428 kJ/mol at 298 K. OH• forms through several key mechanisms, including the direct photodissociation of water molecules under ultraviolet radiation: \mathrm{H_2O + h\nu \rightarrow OH^\bullet + H^\bullet} This process occurs prominently in laboratory settings and the upper atmosphere with vacuum UV light (below 185 nm). In combustion environments, such as flames, OH• is generated via radical chain reactions, for example, through the interaction of hydrogen atoms with oxygen: \mathrm{H + O_2 \rightarrow OH^\bullet + O}. Atmospheric lightning also produces OH• by subjecting air to extreme temperatures (up to 30,000 K), causing thermal dissociation of water vapor and molecular oxygen into reactive species including the radical. Detection of OH• relies on its spectroscopic signatures, with ultraviolet absorption being a primary method due to the strong A ^2\Sigma^+ \leftarrow X ^2\Pi transition, featuring a prominent absorption band at 281 nm. Electron spin resonance (ESR) further confirms its presence by detecting the signal, often enhanced through spin-trapping agents like DMPO for sensitivity in complex matrices. The exhibits a short lifetime in air—on the order of 1 second in the —owing to its rapid reactions with trace gases such as and CH₄, limiting its persistence and necessitating continuous production for sustained atmospheric roles.

Chemical Behavior and Reactions

The hydroxyl radical (OH•) exhibits exceptionally high reactivity due to its , making it one of the most potent oxidizing agents in both atmospheric and biological environments. In the gas phase, OH• primarily undergoes reactions with saturated , where it abstracts a to form and an alkyl ; typical rate constants for these abstractions from alkanes range from approximately 10^{-14} to 10^{-12} cm³ molecule⁻¹ s⁻¹ at 298 K, depending on the specific hydrocarbon and site of . For example, the reaction with proceeds with k = (6.4 ± 0.7) × 10^{-15} cm³ molecule⁻¹ s⁻¹ at 298 K, illustrating the radical's efficiency in initiating oxidation chains. Additionally, OH• adds rapidly to carbon-carbon double bonds in alkenes, forming β-hydroxyalkyl radicals; this addition is diffusion-controlled with rate constants typically (2–8) × 10^{-11} cm³ molecule⁻¹ s⁻¹ at 298 K for simple alkenes, significantly faster than abstraction pathways for unsaturated compounds. These reactions enable OH• to propagate chain mechanisms critical to environmental processes. In tropospheric chemistry, the reaction of OH• with (CO) plays a key role in formation under NOx-rich conditions: \ce{OH^\bullet + CO ->[M] HOCO^\bullet -> CO2 + H^\bullet} followed by H• + O₂ → HO₂• and subsequent cycling that net produces O₃; this pathway accounts for a substantial fraction of global OH• consumption and influences urban air quality. In combustion environments, OH• sustains propagation by abstracting from molecules and participating in branching cycles, with concentrations peaking during ignition to accelerate oxidation rates. OH• also serves as a primary atmospheric scavenger, oxidizing volatile organic compounds (VOCs) and other pollutants to prevent their accumulation; it reacts with over 90% of emitted VOCs, converting them into less harmful products like CO₂ and , thus maintaining the troposphere's self-cleansing capacity. In biological contexts, OH• generated via Fenton chemistry induces by attacking cellular components, notably causing DNA strand breaks and base modifications such as formation, which can lead to mutations if unrepaired. This reactivity implicates OH• in aging and diseases like cancer, where elevated levels overwhelm defenses.

Astronomical Observations

In Earth's Atmosphere and Airglow

The hydroxyl radical (OH) plays a significant role in the chemistry of Earth's , where its concentrations exhibit a peak of approximately 10⁶ molecules cm⁻³ at altitudes between 40 and 50 km. This layer corresponds to the upper , where OH acts as a key oxidant influencing the budgets of and other trace gases. The radical's abundance in this region is governed by photochemical production and loss processes, with concentrations decreasing toward lower altitudes due to reduced solar radiation and increasing air density. Observations indicate that these peak values are typical during daytime conditions, reflecting the radical's sensitivity to flux. In the context of , the contributes to the nightside through the excitation mechanism involving the reaction H + O₃ → * + O₂, which produces vibrationally excited * molecules. These excited states subsequently emit light in the near-infrared as they relax, forming the characteristic Meinel bands spanning wavelengths from 700 to 900 nm. The Meinel bands arise from transitions between vibrational levels in the ground electronic state of , with the emission layer peaking around 85-90 km altitude in the , slightly above the stratospheric OH maximum. This serves as a diagnostic tool for upper atmospheric dynamics and temperatures, as the intensity and rotational structure of the bands correlate with local conditions. The diurnal cycle of stratospheric OH is pronounced, driven primarily by solar-driven . During daytime, the main production pathway is the reaction of electronically excited oxygen atoms with : O(¹D) + H₂O → 2 OH, initiated by photolysis under ultraviolet radiation. At night, OH concentrations decline sharply due to the absence of this photolytic source, with residual levels sustained by minor recombination processes involving HO₂ and other radicals. This cycle results in OH levels varying by orders of magnitude between day and night, underscoring its role as a transient atmospheric cleanser. Satellite-based monitoring has provided critical insights into stratospheric OH distributions and their links to . The Microwave Limb Sounder (MLS) instrument on the satellite, launched in 2004, measures OH along with and other species using microwave emissions, enabling global profiles from the upper to the . Data from MLS reveal correlations between elevated OH and regions of ozone loss, particularly in the presence of radicals from human-emitted chlorofluorocarbons, highlighting OH's catalytic role in ozone destruction cycles. These observations, spanning over two decades, support assessments of stratospheric recovery and the impacts of atmospheric variability.

On Lunar and Planetary Surfaces

The presence of hydroxy groups (OH) on the lunar surface has been detected through and impact experiments, revealing their incorporation into the regolith as a result of interactions with particles. The Mineralogy Mapper (M³) instrument aboard India's spacecraft, launched in 2008, identified widespread absorptions at approximately 3 μm indicative of OH and H₂O in the lunar , particularly at high latitudes and in permanently shadowed regions. This detection was complemented by NASA's Lunar Crater Observation and Sensing Satellite (LCROSS) mission in 2009, which impacted the Cabeus crater and analyzed the ejecta plume, confirming the presence of water molecules (including OH components) at concentrations up to 5.6% by weight in the . These findings attribute the OH formation primarily to the of oxygen atoms in the regolith by hydrogen ions. On Mars, spectroscopic observations from the orbiter, operational since 2003, have revealed seasonal variations in OH signatures within the polar caps, associated with processes in the underlying . The Observatoire pour la Minéralogie, l'Eau, les Glaces et l'Activité (OMEGA) instrument detected broad 3 μm absorptions consistent with OH-bearing phases in the near the south polar cap, with intensities varying seasonally due to temperature-driven exchange between surface ice and subsurface hydrated minerals. These observations link the transient OH enrichment to the and recondensation of from layers during polar summer, influencing the state of phyllosilicates and other minerals in the cap margins. In the Venusian atmosphere, transient OH has been observed in the upper cloud layers (around 65–70 km altitude) through , arising from photochemical reactions involving (SO₂) and (H₂O). The Visible and Infrared Thermal Imaging Spectrometer (VIRTIS) on the mission (2005–2014) detected OH nightglow emissions at 2.8 μm and 1.46 μm on the planet's nightside, indicating short-lived OH radicals formed via oxidation processes in the . These emissions are linked to reactions such as the interaction of SO₂ with H₂O under irradiation, contributing to the dynamic sulfur-water chemistry that sustains the upper cloud haze. The formation of OH on airless planetary bodies like the primarily occurs through and driven by exposure. involves the embedding of protons into the top few nanometers of grains, where they react with lattice oxygen to form OH bonds, as demonstrated in simulations using H⁺ and D⁺ beams on lunar analogs. complements this by ejecting surface atoms and molecules, including nascent OH, through momentum transfer from incident , though it also leads to volatile loss; this process is particularly relevant on bodies lacking atmospheres, where maturity influences OH retention.

In Exoplanetary Atmospheres

The presence of the (OH) in exoplanetary atmospheres has been detected primarily in ultra-hot s, where intense stellar irradiation dissociates into OH and other species. These detections provide insights into high-temperature and dynamics. The first confirmed observation occurred in the dayside atmosphere of WASP-33b, an ultra-hot Jupiter with equilibrium temperatures exceeding 2700 K, using high-dispersion with the Subaru/IRD instrument. This revealed OH emission lines at wavelengths around 1-2 μm, indicating thermal of H₂O in the upper atmosphere. Subsequent observations have identified OH absorption features during planetary , employing transmission to probe the atmospheric composition along the line of sight. For instance, in , ground-based high-resolution during transit detected OH absorption with a of 6.1, particularly in the near-infrared, confirming its role as a tracer of in hot Jupiter atmospheres. The () has contributed to related studies by detecting in hot Jupiters like HD 189733b since the 2010s, where models suggest OH formation via H₂O dissociation under UV irradiation, though direct OH lines remain elusive at HST's resolution. The (JWST), operational since 2021, enhances these capabilities with superior near-infrared sensitivity and resolution for transmission spectroscopy, targeting absorption features at 1-2 μm during transits of hot s. While JWST has yet to report direct OH detections as of 2025, its observations of -bearing atmospheres in similar systems enable refined modeling of OH production. Recent ground-based advances include the 2025 detection of OH emission in the dayside of , the hottest known ultra-hot (T_eq ≈ 4600 K), using the SPIRou spectrograph on the Canada-France-Hawaii Telescope, with a of 5.3, further validating OH as a ubiquitous feature in dissociated layers. These signatures imply active driven by stellar EUV radiation, where breaks down into radicals, which can further photolyze to release and oxygen, facilitating hydrodynamic escape and atmospheric mass loss. In ultra-hot Jupiters, such processes sculpt the upper atmosphere, potentially enriching it with oxygen from water dissociation. Although not a direct , presence in cooler, habitable-zone exoplanets could indirectly signal water photolysis and oxygen accumulation, aiding assessments of atmospheric and potential , akin to processes observed in solar system giant planets.

References

  1. [1]
    CH105: Chapter 9 - Organic Compounds of Oxygen - Chemistry
    For example, an alcohol is an organic compound with a hydroxyl (-OH) functional group on an aliphatic carbon atom. Because -OH is the functional group of all ...
  2. [2]
    Alcohol Reactivity - MSU chemistry
    The functional group of the alcohols is the hydroxyl group, –OH. Unlike the alkyl halides, this group has two reactive covalent bonds, the C–O bond and the O–H ...Missing: definition | Show results with:definition
  3. [3]
  4. [4]
    [PDF] Chem101 General Chemistry
    - This is because the hydroxy group can hydrogen bond to water. - It can serve as a donor in one hydrogen bond, and an acceptor in two hydrogen bonds. O. H. H.
  5. [5]
    Optional Supplementary Chapter: Organic Chemistry – Astrobiology
    An oxygen attached to a hydrogen is called a “hydroxyl” group and is extremely important to biochemistry. The simplest hydroxyl containing organic molecules ...
  6. [6]
    Chemoselective Hydroxyl Group Transformation: An Elusive Target
    The hydroxyl group is present in many biologically significant molecules including ~65% of known natural products,9 ~40% of drugs,9 several classes of lipids ...
  7. [7]
    Functional Groups - FSU Chemistry & Biochemistry
    Select the longest continuous carbon chain, containing the hydroxyl group, and derive the parent name by replacing the -e ending with -ol. Number the carbon ...
  8. [8]
    alcohols (A00204) - IUPAC Gold Book
    Compounds in which a hydroxy group, − OH , is attached to a saturated carbon atom R A 3 COH . The term 'hydroxyl' refers to the radical species, HO A ∙ .
  9. [9]
    [PDF] Nomenclature of Inorganic Chemistry | IUPAC
    Chemical nomenclature must evolve to reflect the needs of the community that makes use of it. In particular, nomenclature must be created to describe new ...
  10. [10]
    What is the difference between hydroxyl, hydroxide and hydroxy?
    Sep 3, 2016 · Hydroxyl radical- {hydro- + oxy(gen) + -yl + radical} - The hydroxyl radical, •OH, is the neutral form of the hydroxide ion (OH-1).
  11. [11]
    Revolutionary Instruments: Lavoisier's Tools as Objets d'Art
    Apr 7, 2008 · Lavoisier created a combustion chamber by inverting a bell jar in a basin of mercury and withdrawing part of the air so that the mercury level ...
  12. [12]
    [PDF] Brief Guide to the Nomenclature of Organic Chemistry - IUPAC
    The International Union of Pure and Applied Chemistry (IUPAC) provides recommendations on many aspects of nomenclature.
  13. [13]
    [PDF] Brief Guide to the Nomenclature of Inorganic Chemistry | IUPAC
    If all hydroxy groups in an oxoacid are replaced, the compound is no longer an acid and is not named as such, but will have a traditional functional class ...
  14. [14]
    Blue Book P-60-65 - IUPAC nomenclature
    Alcohols, phenols, 'enols' and 'ynols' are recognized as important classes of hydroxy compounds. The category is extended so as to include compounds having one ...
  15. [15]
    List of experimental bond lengths for bond type rOH
    List of experimental bond lengths for bond type rOH ; rOH, OH, Hydroxyl radical, 0.970, re ; rOH, H2SO · Sulfuric acid, 0.970.
  16. [16]
    [PDF] Bond Length, Bond Strength & Acidity
    C O. C O. 1.42. 1.22 sp2–sp2. Bond Lengths (Å). • Multiple Bonding: Bond length depends strongly on bond order (length: single > double > triple). • Bond ...
  17. [17]
    1.10: Hybridization of Nitrogen, Oxygen, Phosphorus and Sulfur
    Oct 27, 2023 · The oxygen is sp3 hybridized which means that it has four sp3 hybrid orbitals. One of the sp3 hybridized orbitals overlap with s orbitals from a ...
  18. [18]
    22.3: Names and Properties of Phenols - Chemistry LibreTexts
    Aug 26, 2020 · It was proposed that resonance delocalization of an oxygen non-bonded electron pair into the pi-electron system of the aromatic ring was ...
  19. [19]
    Infrared Spectrometry - MSU chemistry
    The O-H stretching absorption of the hydroxyl group is sensitive to hydrogen bonding. In the gas phase and in dilute CCl4 solution (0.01 M) small to moderate ...Missing: hydroxy | Show results with:hydroxy
  20. [20]
    17.11: Spectroscopy of Alcohols and Phenols - Chemistry LibreTexts
    Sep 24, 2022 · Alcohols and phenols show a broad O-H stretch in IR (3300-3400 cm-1 for alcohols, ~3500 cm-1 for phenols). NMR shows a characteristic -OH ...
  21. [21]
    WebElements Periodic Table » Oxygen » electronegativity
    The first scale of electronegativity was developed by Linus Pauling and on his scale oxygen has a value of 3.44 on a scale running from from about 0.7 (an ...
  22. [22]
  23. [23]
    Estimating the Hydrogen Bond Energy - ACS Publications
    For water, a CCSD(T) interaction energy of 20.8 kJ/mol was reported, (69) for the HF dimer one of about 19.1 kJ/mol (70) to 19.4 kJ/mol was found.
  24. [24]
    Liquids and Gases - Boiling Points - The Engineering ToolBox
    Boiling temperatures for common liquids and gases - acetone, butane, propane and more. ; Air, -191 ; Alcohol - ethyl (grain, ethanol) C2H5OH, 79 ; Alcohol - allyl ...
  25. [25]
  26. [26]
    17.2: Properties of Alcohols and Phenols - Chemistry LibreTexts
    Feb 9, 2025 · Alcohols and water have the ability to form hydrogen bonds with one another which tends to make the two liquids miscible. Small alcohols are ...
  27. [27]
    13.5: Acidity of Alcohols and Phenols - Chemistry LibreTexts
    May 30, 2020 · Phenols are weakly acidic (pKa = 10) because of their ... Alcohols are considered neutral with pKa values similar to water (pKa = 14). The
  28. [28]
    7.4: Acid-base Properties of Phenols - Chemistry LibreTexts
    Jul 20, 2022 · Phenol has a pKa of 10 while 4-methoxyphenol has a pKa of 10.6. Notice that the methoxy group increases the pKa of the phenol group - it makes ...
  29. [29]
    [PDF] values to know 1. Protonated carbonyl pKa
    Approximate pKa chart of the functional groups: values to know. 1. Protonated carbonyl pKa = -7 Other important pKa's. 2. Protonated alcohol or ether pKa = -2 ...
  30. [30]
    [PDF] pKa values
    A ⊕-charged oxygen is called an oxonium ion. The pKa value is typically between -6 and 0. Do not confuse the protonated amine with the neutral amine.
  31. [31]
    17.2: Properties of Alcohols and Phenols - Chemistry LibreTexts
    Nov 10, 2024 · Compare, for instance, the acidities of ethanol (pKa = 16) and 2,2,2-trifluoroethanol (pKa = 12.43), or of tert-butyl alcohol (pKa = 18) and ...
  32. [32]
    4.5: Factors affecting the SN2 Reaction - Chemistry LibreTexts
    May 24, 2021 · For example, thiols (R-SH) are more nucleophilic than alcohols (R-OH) because oxygen has a higher electronegativity than sulfur. Also ...
  33. [33]
    Acidities of Alcohols - Chemistry LibreTexts
    Jan 22, 2023 · Alcohols are very weak Brønsted acids with pKa values generally in the range of 15 - 20. Because the hydroxyl proton is the most electrophilic ...
  34. [34]
    [PDF] Synthesis and Structure of Alcohols
    Alcohols are usually classified as primary, secondary and tertiary. Alcohols with the hydroxyl bound directly to an aromatic (benzene) ring are called phenols.
  35. [35]
    [PDF] Review: Step by Step from Organic Chemistry to Biochemistry
    STEP 7. Polyols: variations on a theme. If a carbon compound has two -OH functional groups, it is called a diol. If it has three, it is called a triol.
  36. [36]
    Chemical Reactivity - MSU chemistry
    Phenol serves as a model for the enol tautomer of cyclohexanone, the aromaticity of the benzene ring stabilizing the hydroxyl form. The enhanced acidity of ...
  37. [37]
    [PDF] Reactions of Alcohols
    Alcohol dehydration usually occurs via the E1 mechanism. The first step is the exothermic protonation of the hydroxyl, followed by the slow, endothermic, rate ...Missing: hydroxy | Show results with:hydroxy
  38. [38]
    Calcium Hydroxide | Ca(OH)2 | CID 6093208 - PubChem
    Pure calcium hydroxide paste has a high pH (approximately 12.5-12.8) and is classified chemically as a strong base. Its main actions are achieved through the ...
  39. [39]
    Influence of soluble oligomeric aluminum on precipitation in the Al ...
    Nov 14, 2020 · Precipitation in this system entails a transition from predominantly tetrahedrally coordinated aluminate (Al(OH)4-) species in solution to ...
  40. [40]
    [PDF] NUREG/CR-6915, "Aluminum Chemistry in a Prototypical Post-Loss ...
    Aluminate ions (Al(OH) 4 ) are the only stable form of aluminum in an alkaline solution. The solubility is a function of solid hydroxide phases and ...
  41. [41]
    O-GlcNAc in cancer biology - PubMed
    O-linked β-N-actylglucosamine (O-GlcNAc) is a carbohydrate post-translational modification on hydroxyl groups of serine and/or threonine residues of cytosolicMissing: hydroxy | Show results with:hydroxy
  42. [42]
    3-hydroxy-4-oxoquinaldine 2,4-dioxygenase: effect of the basicity of ...
    Mar 21, 2014 · We have investigated the role of the active site His-251/Asp-126 dyad, proposed to be involved in substrate hydroxyl group deprotonation, a critical ...
  43. [43]
    4.1: Carbohydrates - Biology LibreTexts
    Apr 27, 2019 · Because carbohydrates have many hydroxyl groups associated with the molecule, they are therefore excellent H-bond donors and acceptors.
  44. [44]
    Antioxidative and Anti-Inflammatory Activity of Ascorbic Acid - PMC
    Moreover, due to the proximity of the hydroxyl and carbonyl groups, ascorbic acid is an ideal hydrogen or electron donor, which makes it the cofactor of many ...3.2. Ros Scavenging By... · 3.3. Ascorbic Acid... · 4. Ascorbic Acid And...Missing: hydroxy | Show results with:hydroxy
  45. [45]
  46. [46]
    Glycosylation as a Main Regulator of Growth and Death Factor ... - NIH
    In this review, we provide a thorough picture of the mechanisms bywhich glycosylation affects the activity of growth and death factor receptors in normal and ...3.2. Galectin Binding · 4. How Glycosylation... · 4.1. Receptors Of The Erbb...Missing: hydroxy | Show results with:hydroxy
  47. [47]
    [PDF] 3.8 ADDITION OF WATER TO AN ALKENE H or enzyme + H-O-H ...
    Much of the industrial (as opposed to beverage) ethanol in this country is actually made by this process. The ethene is a byproduct of the petroleum refineries ...
  48. [48]
    Hydration of Alkenes With Aqueous Acid - Master Organic Chemistry
    Sep 15, 2023 · Hydration of Alkenes to Give Alcohols. When alkenes are treated with aqueous acid (H3O+) they can be converted to alcohols.
  49. [49]
    Alkenes - reaction with sulfuric acid, steam & synthesis of alcohols
    Giving ethanol - this is a synthetic method of making alcohols from alkenes from cracking oil hydrocarbons. The catalyst is phosphoric acid and the reaction ...
  50. [50]
    10.6 : Alcohols from Carbonyl Compounds: Reduction - JoVE
    The three major pathways to reduce carbonyls to alcohols are catalytic hydrogenation, hydride reduction, and borane reduction.Missing: NaBH4 | Show results with:NaBH4
  51. [51]
    Fatty Alcohol - an overview | ScienceDirect Topics
    Fatty alcohols have numerous commercial applications including their use as lubricants, surfactants, solvents, emulsifiers, plasticizers, emollients, thickeners ...
  52. [52]
    Maps and Data - Global Ethanol Production by Country or Region
    The United States is the world's largest producer of ethanol, having produced over 15 billion gallons in 2021 and 2022. Together, the United States and Brazil ...
  53. [53]
    Transport biofuels – Renewables 2020 – Analysis - IEA
    Global ethanol production to drop almost 15% in 2020, mainly due to lower output in Brazil and the United States. Fuel ethanol production reached 115 billion L ...
  54. [54]
    Food, Cosmetics & Explosives – The Versatility of Glycerol
    May 25, 2014 · Glycerol is used in food as a sweetener and humectant, in cosmetics as a smoothing agent, and as a precursor to explosives.
  55. [55]
    Glycerol: A Review on Properties, Industrial and Pharmaceutical ...
    Apr 10, 2023 · Industrial Applications​​ Glycerol is widely used in the manufacture of drugs, medicine and pharmaceuticals for the purpose of dissolving drugs, ...
  56. [56]
    Fatty Alcohols 101: Uses, benefits, and why they matter in everyday ...
    May 7, 2025 · Fatty alcohols are used in skincare, haircare , detergents, pharmaceuticals, and industrial applications. They act as emulsifiers, surfactants, ...What are fatty alcohols? · Fatty alcohols vs. fatty acids...
  57. [57]
    Two-Step Synthesis of Paracetamol (Acetaminophen), a Practical ...
    Sep 15, 2023 · Paracetamol has been reported to be synthesized by various routes, including Beckmann rearrangement of 4-hydroxyacetophenone oxime with a ...Introduction · Results and Discussion · Supporting Information · Author Information
  58. [58]
    (PDF) Phenol Biodegradation: A review - ResearchGate
    Aug 10, 2025 · The release of phenolic compounds in the effluents of petrochemical, textile and coal industry has resulted in contamination of receiving ...
  59. [59]
    A comprehensive review on eco-toxicity and biodegradation of ...
    Sometimes, the phenol biodegradation process produces more toxic intermediate compounds which may exaggerate the toxicity of living cells and tissue. Inside ...
  60. [60]
  61. [61]
    [PDF] Bond dissociation energies in simple molecules
    Bond dissociation energy values (kcal/mol) and (kj/mol) of simple compounds are tabulated ... of the OH free radical by mass spectrometry, J. Chem. Phys ...
  62. [62]
    [PDF] Kinetics of the Hydroxyl Radical - Standard Reference Data
    H+03→ OH*+02. Reaction rates involving hydroxyl radicals are required for predictions of the vertical distribution of ozone [10] and for the calculation ...<|control11|><|separator|>
  63. [63]
    [PDF] RADICALS IN FLAMES
    Some of the important radicals in hydrogen and hydro- carbon flames are hydrogen atoms (H), hydroxyl radicals (OH), oxygen atoms (0), and methyl radicals (CH3).
  64. [64]
    Electrical Discharges Produce Prodigious Amounts of Hydroxyl and ...
    Apr 29, 2021 · Lightning is known to generate nitrogen oxides and ozone, which indirectly create the hydroxyl radical (OH), the atmosphere's primary oxidant.
  65. [65]
    Trends in global tropospheric hydroxyl radical and methane lifetime ...
    Nov 5, 2020 · 2.2 Inferred OH from measurements. Tropospheric OH has a chemical lifetime of less than a second or so, reflecting its high reactivity and ...
  66. [66]
    Relative rate constants for reaction of the hydroxyl radical with a ...
    Relative rate constants for reaction of the hydroxyl radical with a series of alkanes, alkenes, and aromatic hydrocarbons | The Journal of Physical Chemistry.
  67. [67]
    II Reactions of Alkenes with the OH Radical - Oxford Academic
    Oct 31, 2023 · The chemical reactivity of the alkenes toward OH, 03, N03, and 0(3P) is much greater than that for alkanes. This is illustrated for the ...Missing: addition | Show results with:addition<|control11|><|separator|>
  68. [68]
    7.4.5 The Hydroxyl Radical - AR4 WGI Chapter 7 - IPCC
    The primary source of tropospheric OH is a pair of reactions starting with the photodissociation of ozone by solar UV radiation. Additionally, in the remote, ...<|control11|><|separator|>
  69. [69]
    Re-assessing hydroxyl radical chemistry in the atmosphere - Nature
    Apr 26, 2025 · Globally, the hydroxyl radical (OH) is the primary oxidant in the daytime troposphere where it controls the removal of greenhouse gases such as ...
  70. [70]
    Hydroxyl radical is a significant player in oxidative DNA damage in ...
    Jun 15, 2021 · Recent publications have suggested that oxidative DNA damage mediated by hydroxyl radical ( • OH) is unimportant in vivo, and that carbonate anion radical (CO ...Missing: review | Show results with:review
  71. [71]
    Oxidative Stress: Harms and Benefits for Human Health - PMC - NIH
    For instance, an excess of hydroxyl radical and peroxynitrite can cause lipid peroxidation, thus damaging cell membranes and lipoproteins.Oxidative Stress: Harms And... · 2. Oxidants And Free Radical... · 6. Prooxidant Agents In...
  72. [72]
    Hydroxyl Radical - an overview | ScienceDirect Topics
    Hydroxyl radicals are highly reactive species that attack most of the organic molecules. They are highly oxidizing in nature which is attributed to their ...<|control11|><|separator|>
  73. [73]
    Hydroxyl in the stratosphere and mesosphere – Part 1: Diurnal ...
    Feb 2, 2011 · The primary driver for OH diurnal variations is the ultraviolet actinic flux that initiates the photochemical production of reactive hydrogen ...Missing: radical | Show results with:radical
  74. [74]
    Hydroxyl airglow observations for investigating atmospheric dynamics
    Jan 27, 2023 · Here, we briefly summarize why an OH* airglow layer exists, how atmospheric dynamics influence it and how temperature can be derived from OH* ...
  75. [75]
    [PDF] SABER observations of the OH Meinel airglow variability near the ...
    In this paper we compare satellite observations of OH Meinel emissions to simulations from a 3-dimensional chemical-dynamical model, in order i) to test the ...
  76. [76]
    [PDF] The Earth Observing System Microwave Limb Sounder (EOS MLS ...
    Climate change could possibly delay ozone recovery through stratospheric cooling that can exacerbate some ozone chemical destruction processes, and possibly ...<|control11|><|separator|>
  77. [77]
    The 27 day solar rotational effect on mesospheric nighttime OH and ...
    MLS Aura satellite pressure levels, the corresponding altitudes, and availability for OH and O3 measurements. Furthermore, the impact of large solar proton ...
  78. [78]
    LCROSS - NASA Science
    13, 2009, NASA formally announced that data from LCROSS “indicates that the mission successfully uncovered water … near the Moon's south pole.” Nearly a ...
  79. [79]
    A solar wind-derived water reservoir on the Moon hosted by impact ...
    Mar 27, 2023 · Impact glass beads can store substantial quantities of solar wind-derived water on the Moon and suggest that impact glass may be water reservoirs on other ...
  80. [80]
    Hydration state of the Martian surface as seen by Mars Express ...
    Aug 29, 2007 · This feature is caused by bending and stretching vibrations of adsorbed or structural H2O and/or OH- bound to minerals. ... hydroxyl groups or ...
  81. [81]
    Water in the Martian regolith from OMEGA/Mars Express - Audouard
    Aug 11, 2014 · This reservoir is characterized by the presence of H2O- and OH-bearing phases that produce a broad absorption at a wavelength of ~3 µm in near- ...
  82. [82]
    Sulfur dioxide in the Venus atmosphere: I. Vertical distribution and ...
    Photochemical oxidation efficiently removes SO2 from Venus' upper atmosphere leading to the formation of the sulfuric acid droplets making up the clouds and ...Missing: H2O | Show results with:H2O
  83. [83]
    Hydroxyl (OH) production on airless planetary bodies
    Hydroxyl (OH) production on airless planetary bodies: Evidence from H+/D+ ... Our results emphasize the need to use lunar soils with space-weathered exteriors in ...
  84. [84]
    Solar Wind Sputtering Rates of Small Bodies and Ion Mass ...
    Sep 5, 2017 · Solar wind sputtering of airless bodies ejects abundant secondary ions into the exosphere surrounding the body Sputtered secondary ion ...
  85. [85]
    [2110.11946] Detection of OH in the ultra-hot Jupiter WASP-76b - arXiv
    Oct 22, 2021 · The paper reports the detection of OH in the atmosphere of WASP-76b, an ultra-hot Jupiter, with a peak signal-to-noise ratio of 6.1.Missing: group exoplanet
  86. [86]
    WATER VAPOR IN THE SPECTRUM OF THE EXTRASOLAR ...
    The secondary star HD 189733B is at 7 o'clock with respect to the primary HD 189733A. We oriented HST so as to separate the two stars vertically on the ...Missing: dissociation | Show results with:dissociation
  87. [87]
    Detection of OH and Fe in the dayside atmosphere of the hottest ...
    Here, we present the first detection of the OH molecule in the dayside atmosphere of KELT-9b, based on two thermal emission observations conducted with the ...
  88. [88]
    Atmospheric Escape From Three Terrestrial Planets in the L 98-59 ...
    We find that, regardless of their initial water content, all three planets accumulate significant quantities of oxygen due to efficient water photolysis and ...
  89. [89]
    Atmospheric Beacons of Life from Exoplanets Around G and K Stars
    Nov 2, 2017 · In water-rich exoplanetary atmospheres, photolysis of atmospheric water vapor can also produce an accumulation of oxygen due to the thermal ...