Nitric oxide (NO) is a colorless diatomic gas and free radical with the chemical formula NO, consisting of one nitrogen atom bonded to one oxygen atom and possessing an unpaired electron that confers paramagnetic properties. Beyond biology, NO is a key intermediate in industrial nitric acid production and contributes to atmospheric pollution as a precursor to ozone and acid rain.[1] As a small, uncharged molecule, NO exhibits high diffusivity and solubility in hydrophobic environments such as cell membranes and lipid bilayers, enabling it to traverse biological barriers without the need for transporters.[2] Chemically, NO is relatively unreactive with most biological nucleophiles but rapidly interacts with other radicals like superoxide (O₂⁻•) and molecular oxygen (O₂), as well as transition metals such as iron in hemoproteins, leading to the formation of reactive nitrogen species like nitrite (NO₂⁻), nitrate (NO₃⁻), and peroxynitrite (ONOO⁻).[2] In biological systems, NO serves as a versatile signaling mediator produced endogenously by three isoforms of nitric oxide synthase (NOS) enzymes—neuronal (nNOS), inducible (iNOS), and endothelial (eNOS)—which catalyze the conversion of L-arginine to NO and L-citrulline in the presence of oxygen and cofactors.[3]Physiologically, NO plays pivotal roles in cardiovascular homeostasis, neurotransmission, and immune defense.[3] In the vasculature, eNOS-derived NO activates soluble guanylyl cyclase (sGC) in smooth muscle cells, elevating cyclic guanosine monophosphate (cGMP) levels to induce vasodilation, inhibit platelet aggregation, and regulate blood pressure.[3] As a retrograde neurotransmitter, NO facilitates synaptic plasticity and long-term potentiation in the brain by diffusing across synapses and modulating neuronal excitability through cGMP-dependent pathways.[3] In the immune system, iNOS produces high concentrations of NO (in the micromolar range) in response to inflammatory stimuli, exerting antimicrobial effects by generating nitrosative stress that damages pathogens like bacteria and parasites, while also modulating T-cell responses and apoptosis.[3] The biological activity of NO is concentration-dependent: low levels (nM range) promote cell survival, proliferation, and anti-inflammatory effects, whereas higher levels (μM range) can induce oxidative and nitrosative stress, leading to protein nitration, DNA damage, and cell death.[4]Medically, NO's diverse functions have led to its therapeutic exploitation, particularly in conditions involving vascular and respiratory dysfunction.[3] Inhaled NO is approved for treating persistent pulmonary hypertension in newborns, where it selectively dilates pulmonary vessels to improve oxygenation without systemic hypotension.[3] NO donors such as nitroprusside and nitroglycerin are used clinically for angina and heart failure, while S-nitrosothiols and other NO donors are under investigation for erectile dysfunction by mimicking endogenous NO to enhance vasodilation.[3] Emerging biomedical applications include NO-releasing nanomaterials for wound healing, antibacterial therapies, and targeted cancer treatment, where controlled NO delivery promotes angiogenesis, collagen synthesis, and tumor cell apoptosis while minimizing toxicity.[3] Dysregulated NO production is implicated in diseases like atherosclerosis, neurodegeneration, sepsis, and cancer, underscoring its dual role as both protector and pathogen in humanhealth.[4]
Structure and properties
Molecular structure
Nitric oxide (NO) is a heteronuclear diatomic molecule composed of a single nitrogen atom bonded to a single oxygen atom, forming a stable radical due to the presence of one unpaired electron. According to molecular orbital (MO) theory, the NO molecule has a bond order of 2.5, reflecting significant multiple-bond character intermediate between a double and triple bond. This bond order arises from the arrangement of its 11 valence electrons (five from nitrogen and six from oxygen) across the molecular orbitals formed by the overlap of atomic 2s and 2p orbitals.In the MO diagram for NO, the valence electrons fill the configuration (\sigma_{2s})^2 (\sigma^*_{2s})^2 (\pi_{2p})^4 (\sigma_{2p})^2 (\pi^*_{2p})^1, with eight electrons in bonding orbitals (σ_{2s}^2, π_{2p}^4, σ_{2p}^2) and three in antibonding orbitals (σ^_{2s}^2, π^_{2p}^1). The bond order is calculated as (number of bonding electrons - number of antibonding electrons)/2 = (8 - 3)/2 = 2.5, consistent with the observed bond length of 1.154 Å.[5]The unpaired electron resides in the π^*_{2p} antibonding orbital, which contributes to the molecule's reactivity as a free radical while maintaining overall stability. This electronic structure renders NO paramagnetic, as the unpaired electron imparts a net magnetic moment to the molecule.[6]
Physical properties
Nitric oxide (NO) is a colorless gas at standard temperature and pressure, appearing as a blueliquid when condensed and a bluish-white solid when frozen. It is odorless at low concentrations but develops a pungent, sharp odor at higher levels.[7][8]The molecular weight of NO is 30.01 g/mol, and its density as a gas is 1.34 g/L at standard temperature and pressure (STP). It has a low boiling point of -151.7 °C and a melting point of -163.6 °C, reflecting its weak intermolecular forces. These properties are summarized in the following table:
[7][9][10]NO exhibits limited solubility in water, approximately 4.6 mL of gas per 100 mL of water at 20 °C, due to its nonpolar character despite a small dipole moment. This dipole moment is measured at 0.159 D, oriented from oxygen to nitrogen, which imparts slight polarity to the molecule and influences its interactions in solution. Solubility is notably higher in organic solvents such as ethanol and carbon disulfide.[7][11][12]Thermodynamically, the standard enthalpy of formation (ΔH_f°) for gaseous NO is +90.25 kJ/mol, indicating an endothermic formation process from its elements. The molar heat capacity at constant pressure (C_p) is approximately 29.9 J/mol·K at 298 K, consistent with its diatomic structure and degrees of freedom.[13]
Electronic and spectroscopic properties
Nitric oxide (NO) has a ground electronic state of ^2\Pi, arising from the molecular orbital configuration (\sigma_{2s})^2 (\sigma^*_{2s})^2 (\pi_{2p})^4 (\sigma_{2p})^2 (\pi^*_{2p})^1, where the unpaired electron occupies the antibonding \pi^*_{2p} orbital, resulting in a bond order of 2.5.[14] This open-shell configuration imparts paramagnetic properties to the molecule, with the unpaired electron primarily localized on the nitrogen atom but delocalized to some extent onto oxygen.[15]The ^2\Pi ground state exhibits significant spin-orbit coupling due to the interaction between the orbital angular momentum (\Lambda = 1) and the electronspin (S = 1/2), splitting the state into ^2\Pi_{1/2} and ^2\Pi_{3/2} components separated by the spin-orbit coupling constant A \approx 121.2 cm^{-1} .[16] This coupling constant varies slightly with vibrational level, decreasing from A_0 = 121.17 cm^{-1} for v=0 to lower values in higher vibrational states, as computed from potential energy curves and spectroscopic data.[17] The effects are observable in rotational fine structure, where the splitting influences the Lambda-doubling and overall energy levels of the molecule.A prominent spectroscopic feature is the γ-band system in the ultraviolet region (approximately 200–230 nm), corresponding to the forbidden but observable electronic transition from the ground X^2\Pi state to the excited A^2\Sigma^+ state.[18] This transition, with its (0,0) band origin near 2260 Å, arises from promotion of the unpaired \pi^*_{2p} electron to a Rydberg-like orbital, and its intensity is enhanced by vibronic coupling; it serves as a key signature for NO detection in astrophysical and atmospheric contexts.[19]In the infrared spectrum, the dominant feature is the fundamental vibrational mode, corresponding to the N=O stretching vibration at approximately 1876 cm^{-1} for the v=0 \to v=1 transition in the ground state.[20] This high-frequency mode reflects the strong triple-bond character influenced by the electronic configuration, with rotational structure showing P- and R-branches split by the spin-orbit interaction in the ^2\Pi state.[21]The unpaired electron in NO also produces a characteristic electron paramagnetic resonance (EPR) spectrum, observable in the gas phase or trapped in matrices, with a g-factor near 2.001 and hyperfine splitting due to the ^{14}\mathrm{N} nucleus (nuclear spin I = 1), resulting in a triplet pattern with splitting constant a_N \approx 31.5 G.[15] The spectrum is anisotropic, reflecting the ^2\Pi ground state, and broadens with pressure due to collisions, making EPR a sensitive probe for NO in low concentrations.[22]
History
Early observations
Early chemists in the 18th century began noting peculiar gases, termed "nitrous air," associated with the products of combustion and reactions involving saltpeter (potassium nitrate), though these were not fully isolated or characterized until later.[23]In 1772, Joseph Priestley isolated nitric oxide, which he called "nitrous air," by reacting nitric acid with copper turnings, marking the first preparation of the pure gas.[24] Priestley described its properties in his work Experiments and Observations on Different Kinds of Air, noting its colorless nature and its ability to support combustion weakly compared to common air; he used it as a test for the "goodness" of air, observing that it diminished in volume when mixed with oxygen-rich air due to the formation of red fumes (nitrogen dioxide).[25]By 1800, Humphry Davy identified nitric oxide as a distinct gas separate from mixtures of nitrogen oxides, through systematic experiments distinguishing it from related compounds like nitrous oxide and nitrogen dioxide.[26] Early notes highlighted its toxicity, with Davy reporting that inhalation caused rapid death in animals, faster than in common air but slower than in highly poisonous gases, accompanied by purple-red discoloration of tissues due to interactions with hemoglobin.[26]
Discovery and key developments
In 1901, Wilhelm Ostwald and his collaborators developed the catalytic oxidation of ammonia to nitric oxide as a key step in the industrial production of nitric acid, marking a significant advancement in the scalable synthesis of NO. This process involved passing ammonia gas over a platinum catalyst in the presence of oxygen at high temperatures, yielding NO as the primary product before further oxidation to nitrogen dioxide. The innovation addressed the growing demand for nitrates in fertilizers and explosives, transforming NO from a laboratory curiosity into an industrially viable compound.[27]During the 1970s, nitric oxide gained recognition as a major atmospheric pollutant, particularly in the context of photochemical smog formation following severe air quality episodes in urban areas like Los Angeles. Studies demonstrated that NO, emitted primarily from vehicle exhaust and combustion sources, reacts with sunlight and hydrocarbons to produce ground-level ozone and other secondary pollutants, contributing to haze and health risks. This understanding prompted regulatory actions, including the U.S. Clean Air Act amendments of 1970, which targeted nitrogen oxides (NOx) as criteria pollutants to mitigate smog-related environmental and public health impacts.[28][29]The 1980s and 1990s brought transformative insights into NO's biological roles, culminating in the 1998 Nobel Prize in Physiology or Medicine awarded to Robert F. Furchgott, Louis J. Ignarro, and Ferid Murad for discovering NO as an endogenous signaling molecule in the cardiovascular system. Furchgott identified endothelium-derived relaxing factor (EDRF) in 1980, Ignarro confirmed its identity as NO in 1986, and Murad elucidated its mechanism via cyclic GMP pathways in the 1970s. These findings revolutionized understanding of vasodilation, platelet aggregation, and neurotransmission, shifting NO's perception from toxin to vital biomolecule.[30]In the 2000s, advances in NO donor compounds—such as S-nitrosothiols, diazeniumdiolates, and hybrid nitrates—emerged as promising therapeutic agents for cardiovascular and inflammatory diseases, enabling controlled NO release to mimic endogenous signaling without toxicity. These developments built on Nobel-recognized mechanisms, leading to clinical applications like improved erectile dysfunction treatments and anti-hypertensive drugs. By the 2020s, NO therapies gained attention for COVID-19 management, with inhaled NO demonstrating potential to improve oxygenation in acute respiratory distress syndrome (ARDS) by enhancing pulmonary vasodilation and inhibiting viral replication in preclinical models. Retrospective studies and trials reported reduced ventilation needs in moderate cases, though larger randomized trials are ongoing to confirm efficacy. As of 2025, recent randomized controlled trials have shown that inhaled NO improves respiratory outcomes in mild to severe COVID-19 cases, with benefits including better oxygenation and safety profiles, though further studies are needed for widespread clinical guidelines.[31][32][33][34]
Synthesis
Natural sources
Nitric oxide (NO) is produced endogenously in various organisms through the action of nitric oxide synthase (NOS) enzymes, which catalyze the conversion of L-arginine to NO and L-citrulline in the presence of oxygen and NADPH.[35] In mammals, three main isoforms of NOS exist—neuronal (nNOS), inducible (iNOS), and endothelial (eNOS)—each playing distinct roles in signaling, immune response, and vascular function.[35] Bacteria also possess NOS-like enzymes, such as those identified in Streptomyces species, which share structural similarities with mammalian counterparts and contribute to NO-mediated processes like antibiotic production and stress response.[36] In plants, NO is produced through various non-canonical pathways, such as the reduction of nitrite by nitrate reductase under acidic or hypoxic conditions, supporting processes like root development, pathogen defense, and stress adaptation. No canonical NOS enzymes similar to those in animals have been identified in higher plants.[37]In the environment, NO forms atmospherically during lightning strikes, where high temperatures drive the reaction of nitrogen and oxygen gases: N₂ + O₂ → 2NO.[38] This process contributes significantly to natural NOx (NO + NO₂) production, with global estimates ranging from 2 to 20 Tg N per year, depending on lightning frequency and energy dissipation models.[39] Soil microbial processes, particularly denitrification by bacteria and fungi under anaerobic conditions, release NO as an intermediate when nitrate or nitrite is reduced to gaseous nitrogen products.[40]Minor natural sources include volcanic emissions, where NO is generated through high-temperature reactions in magma or lightning within eruption plumes, contributing approximately 0.02 Tg N per year globally.[41]Ocean photochemistry also produces NO via the photolysis of dissolved nitrite in seawater under sunlight, with rates influenced by pH, temperature, and nitrite concentrations, though this flux remains small compared to terrestrial sources.[42]Overall, natural processes—encompassing biogenic NOS activity, lightning, soil denitrification, and minor geogenic inputs—account for an estimated global NO flux of 10–20 Tg N per year, primarily from soils and atmospheric fixation.[43]
Industrial production
Nitric oxide is primarily produced on an industrial scale as an intermediate in the manufacture of nitric acid via the Ostwald process, which accounts for the vast majority of global production. In this process, ammonia is oxidized over a platinum-rhodium catalyst at temperatures around 850–900°C, following the reaction 4NH₃ + 5O₂ → 4NO + 6H₂O. The resulting process gas contains approximately 10–11% NO, which is then further oxidized to nitrogen dioxide (NO₂) for absorption into water to form nitric acid. This method achieves high conversion rates, with ammonia yields exceeding 95% under optimized conditions, making it the dominant pathway for large-scale NO generation.[44][45]Historically, nitric oxide was produced directly from atmospheric nitrogen and oxygen through the Birkeland-Eyde process, developed in the early 20th century. This electric arc method heats air to over 3,000°C to drive the reaction N₂ + O₂ → 2NO, yielding a gas mixture with about 1–2% NO, which is subsequently converted to nitric acid. Although energy-intensive and largely superseded by the more efficient Ostwald process, it represented a pioneering approach to nitrogen fixation and remains relevant in discussions of plasma-based synthesis.[46]Global production of nitric oxide is tied closely to nitric acid demand, with approximately 69 million metric tons of nitric acid produced annually in 2025, implying a comparable scale of NO as an intermediate in NOx-equivalent forms. Pure NO, however, is manufactured on a much smaller scale, primarily through specialized oxidation or reduction processes, to meet niche demands. In modern applications, high-purity NO serves as a reagent in semiconductor manufacturing, particularly for selective etching processes such as the removal of silicon nitride (SiN) over silicon dioxide (SiO₂) or silicon (Si) using NF₃ plasma, where it enhances fluorine migration and etching selectivity.[47][48]
Laboratory preparation
One common laboratory method for preparing nitric oxide (NO) involves the reduction of dilute nitric acid with copper metal, which produces NO gas along with copper(II) nitrate and water. The balanced equation for this reaction is:$3\mathrm{Cu} + 8\mathrm{HNO_3} \rightarrow 3\mathrm{Cu(NO_3)_2} + 2\mathrm{NO} + 4\mathrm{H_2O}This classic approach is favored for its simplicity and use of readily available reagents, with the reaction typically conducted at room temperature to minimize side products like nitrogen dioxide.Another method entails the thermal decomposition of nitrosyl chloride (NOCl), yielding NO and chlorine gas. The decomposition reaction is:$2\mathrm{NOCl} \rightarrow 2\mathrm{NO} + \mathrm{Cl_2}This process occurs upon heating NOCl, often prepared separately from nitric acid and hydrochloric acid, and is useful for generating pure NO in controlled amounts for subsequent reactions.For spectroscopic studies, NO is frequently generated via gas-phase reactions or from commercial sources and isolated in argon matrices at cryogenic temperatures (typically 10–20 K) to stabilize the molecule and its complexes for infrared or other analyses. These matrix-isolation techniques prevent unwanted dimerization or reactions with oxygen, enabling detailed examination of NO's vibrational spectra and interactions with other species.[49]Purification of laboratory-prepared NO commonly involves trapping the gas in liquid nitrogen (at 77 K) to condense it while removing higher-boiling impurities like water or nitrogen dioxide, followed by fractional distillation under reduced pressure to achieve high purity levels suitable for sensitive experiments.[50]
Chemical reactions
Reactions with small molecules
Nitric oxide (NO) undergoes a third-order gas-phase oxidation reaction with molecular oxygen (O₂) to form nitrogen dioxide (NO₂), according to the equation:$2\text{NO} + \text{O}_2 \rightarrow 2\text{NO}_2This reaction follows the rate law rate = k[NO]²[O₂], where the rate constant k at 300 K is approximately 7.0 × 10³ L² mol⁻² s⁻¹, with an Arrhenius expression k = 1.2 × 10³ exp(530/T) L² mol⁻² s⁻¹ over temperatures from 273 to 600 K.[51] The process is significant in atmospheric chemistry, particularly as a key step in the formation of photochemical smog, where NO emitted from combustion sources oxidizes to NO₂, which then photolyzes to initiate ozone production.NO also reacts with hydrogen (H₂) in a catalytic hydrogenation process to produce nitrogen (N₂) and water (H₂O), represented by:$2\text{NO} + 2\text{H}_2 \rightarrow \text{N}_2 + 2\text{H}_2\text{O}This reduction typically requires metal catalysts such as platinum supported on alumina (Pt/Al₂O₃), with activity increasing with Pt loading and optimal performance in the temperature range of 273 to 373 K. The reaction proceeds via surface adsorption of NO and H₂, leading to stepwise dissociation and recombination, and is studied for applications in exhaust gas treatment to mitigate NOx emissions.In the presence of ozone (O₃), NO rapidly converts to NO₂ and O₂ through the exothermic reaction:\text{NO} + \text{O}_3 \rightarrow \text{NO}_2 + \text{O}_2The rate constant for this elementary reaction is temperature-dependent, k = 3.0 × 10^{-12} exp(−1500 / T) cm³ molecule⁻¹ s⁻¹ over 200–400 K.[52] This process plays a central role in stratospheric and tropospheric ozone depletion cycles, where NO acts as a catalyst by regenerating via subsequent photolysis of NO₂, resulting in net O₃ destruction without consuming NO.The formation of NO from nitrogen (N₂) and oxygen (O₂) is governed by the endothermic equilibrium:\text{N}_2 + \text{O}_2 \rightleftharpoons 2\text{NO}with a standard enthalpy change ΔH° = +180 kJ mol⁻¹ (based on ΔH_f° = +90 kJ mol⁻¹ for NO).[13] This equilibrium favors NO production at high temperatures, as seen in combustion processes like the Zeldovich mechanism, but the yield remains low due to the large positive ΔH and high activation energy, typically requiring temperatures above 2000 K for significant conversion.[13]
Reactions in organic chemistry
Nitric oxide participates in nitrosation reactions with alcohols in non-aqueous solvents, yielding alkyl nitrites (RONO) as key intermediates in organic synthesis. This method involves bubbling gaseous NO through a solution of the alcohol under aerobic conditions, where oxygen facilitates the formation of reactive species that enable the O-nitrosation. For instance, primary and secondary alcohols such as ethanol and isopropanol react efficiently to produce ethyl nitrite and isopropyl nitrite in yields exceeding 80%, offering a mild alternative to traditional syntheses using nitrous acid or nitrosyl chloride.[53]As a free radical, nitric oxide adds to carbon-carbon double bonds of alkenes, initiating radicalchain processes that lead to nitroso compounds. The addition typically forms β-nitrosoalkyl radicals, which can dimerize or undergo further oxidation to yield stable C-nitroso products, often in the presence of oxygen to prevent reversion. A representative example is the reaction with cyclohexene, producing bis(1-nitroso-2-nitrocyclohexane) through initial NO• addition followed by trapping of the adduct radical by NO2 derived from NO autoxidation; crystal structures confirm the trans configuration of the nitro and nitroso groups. This reactivity is particularly useful for synthesizing α-nitrosonitro compounds from simple alkenes, though yields are moderate (20-50%) due to competing dimerization.[54][55]Nitric oxide appears as a byproduct in certain diazotization processes of aromatic amines, arising from side reactions during nitrous acid generation, and can be harnessed for sustainable synthetic routes. In conventional diazotization with NaNO2 and acid, excess nitrite or improper pH can lead to NO evolution via decomposition pathways, complicating product isolation. Conversely, low-concentration NO streams from industrial denitrification can be directly coupled with aniline diazotization in aqueous media, achieving up to 95% conversion to diazonium salts while recycling NOx emissions; this approach minimizes waste and uses NO as an effective nitrosating agent under mild conditions (0-5°C).[56]In C-H nitration reactions, nitric oxide acts as an oxidant precursor, enabling selective aromatic nitration through derived reactive nitrogen species. Under aerobic conditions, NO reacts with O2 to form NO2, which abstracts a hydrogen from activated C-H bonds (e.g., in phenols or tyrosine derivatives), followed by nitro group incorporation. For example, treatment of 4-methylphenol with NO/O2 affords 2-nitro-4-methylphenol in 60-70% yield, highlighting NO's role in generating electrophilic nitrating agents without strong acids; this method is valuable for late-stage functionalization in complex organic molecules.
Coordination and complex formation
Nitric oxide (NO) acts as an ambidentate ligand in transition metal coordination chemistry, primarily binding through the nitrogen atom to form nitrosyl complexes, though rare O-bound isomers exist. The M–N–O linkage can adopt a linear geometry (bond angle ≈ 180°) characteristic of the electrophilic NO⁺ form or a bent geometry (bond angle ≈ 120–150°) indicative of the nucleophilic NO⁻ form, with the neutral NO• radical exhibiting intermediate behavior depending on the metal's electronic properties.To classify these electronic variations without assigning ambiguous formal oxidation states, the Enemark–Feltham notation designates nitrosyl complexes as {M(NO)x}n, where M is the metal, x is the number of NO ligands, and n represents the total number of electrons in the metal d orbitals plus the π* orbitals of the NO ligands (treating each NO as contributing one π* electron). For example, a neutral NO ligand in {NO}0 corresponds to the radical form, while {NO}1 denotes NO⁺ and {NO}-1 denotes NO⁻. This formalism, introduced in 1974, facilitates understanding of bonding and reactivity across diverse metal nitrosyls.The bonding in metal nitrosyl complexes involves σ-donation from the NO 5σ orbital to an empty metal d orbital, augmented by π-backbonding from filled metal d orbitals into the antibonding π* orbitals of NO, which weakens the N–O bond and influences its redox behavior. Depending on the metal's oxidation state and ligand field, NO can function as NO⁺ (strong σ-donor, weak π-acceptor in linear mode), NO⁻ (strong π-donor in bent mode), or NO• (balanced donor-acceptor with radical character). These interactions lead to delocalized electronic structures, often requiring spectroscopic and computational methods for precise description.Representative examples include sodium nitroprusside, Na2[Fe(CN)5(NO)]·2H2O, where the [Fe(NO)(CN)5]2– anion features a linear Fe–N–O bond and is classified as {FeNO}6, consistent with an FeII(NO⁺) description; this complex serves as a photolabile NO donor in medical applications. Another notable complex is trichloronitrosylruthenium, Ru(NO)Cl3, a {RuNO}6 species with a linear Ru–N–O linkage, employed for its vasodilatory effects due to controlled NO release.[57]Metal nitrosyl complexes play key roles in catalysis, particularly in the selective catalytic reduction (SCR) of NO in automotive exhaust systems, where surface-bound nitrosyl intermediates on metals like copper or iron facilitate NO dimerization to N2O or further reduction to N2. For instance, in Cu-zeolite catalysts, side-on {CuNO}2 species act as pivotal intermediates in ammonia-SCR processes, enabling efficient NOx abatement under lean-burn conditions.[58]
Detection methods
Analytical techniques
Chemiluminescence detection is a widely used physicochemical method for quantifying nitric oxide (NO) in gaseous samples, particularly in air quality monitoring. The technique relies on the reaction of NO with ozone (O₃), where NO combines with O₃ to form excited-state nitrogen dioxide (NO₂*), which subsequently emits light upon relaxation to the ground state:\text{NO} + \text{O}_3 \rightarrow \text{NO}_2^* + \text{O}_2 \rightarrow \text{NO}_2 + h\nuThe emitted light intensity, typically in the red to near-infrared range (>600 nm), is measured by a photomultiplier tube and correlates linearly with NO concentration. This method serves as the reference standard for ambient NO and NO₂ measurements, offering high sensitivity (detection limits around 1 ppb) and specificity when combined with catalytic converters to distinguish NO from NO₂.[59]Electrochemical sensors provide real-time amperometric detection of NO in both gaseous and aqueous solutions, operating through the electrooxidation of NO at electrode surfaces. These sensors typically employ platinum (Pt) working electrodes polarized at potentials of 0.6–0.9 V versus Ag/AgCl, facilitating a three-electron oxidation process: NO → NO⁺ → NO₂ → NO₃⁻. Platinized Pt electrodes, coated with platinum black to increase surface area, enhance sensitivity by up to 10-fold, achieving detection limits as low as 83 pM in physiological solutions. Permselective membranes, such as Nafion or fluorinated xerogels, are often applied to minimize interferences from species like nitrite or ascorbic acid, enabling selective NO quantification in complex matrices.[60]Mass spectrometry offers direct identification and quantification of NO in gases and solutions via detection of the NO⁺ ion at m/z 30. Membrane inlet mass spectrometry (MIMS) is particularly effective for aqueous samples, where a semipermeable membrane allows NO to diffuse into the ion source of a quadrupole mass spectrometer, ionizing it to NO⁺ for selective monitoring. This approach provides high specificity, with a lower detection limit of 10 nM and a linear range up to 50 μM, making it suitable for precise measurements in controlled laboratory settings without derivatization.[61]The Griess reaction serves as an indirect colorimetric assay for NO in solutions by quantifying nitrite (NO₂⁻), a stable oxidation product of NO. In this method, nitrite reacts with sulfanilamide under acidic conditions to form a diazonium salt, which then couples with N-(1-naphthyl)ethylenediamine to produce a purple azo dye absorbing at 540 nm. The assay detects nitrite concentrations as low as 0.5 μM, providing a proxy for NO levels after sample oxidation or incubation, though it requires careful control of reaction conditions to ensure accuracy in aqueous media.[62]
Biological and environmental assays
In biological systems, nitric oxide (NO•) detection often relies on fluorescent probes that enable real-time imaging in living cells. The diamino fluorescein derivative DAF-FM, a cell-permeable indicator, is widely used for this purpose; it diffuses into cells, where esterases convert it to the impermeable DAF-FM, which then reacts with NO-derived nitrosating species (such as N2O3) to form a highly fluorescent triazolo-fluorescein product, allowing visualization of intracellular NO• production via fluorescence microscopy. This probe exhibits high sensitivity, detecting NO• at nanomolar concentrations, and has been applied in studies of neuronal and endothelial cells to monitor NO• signaling dynamics.Electron spin resonance (ESR) spectroscopy provides a direct method for detecting the NO• free radical in biological tissues by trapping it with iron-dithiocarbamate complexes, such as Fe²⁺-MGD, forming a stable mononitrosyl-iron complex that produces a characteristic three-line ESR spectrum.[63] This technique is particularly valuable for quantifying NO• in complex matrices like blood or organ homogenates, where it distinguishes NO• from other reactive nitrogen species, and has been employed to assess NO• levels in ischemic tissues or during inflammation.[64]Indirect biosensors for NO• in biological contexts often utilize NO• scavengers or inhibitors of nitric oxide synthase (NOS). For instance, oxyhemoglobin assays measure NO• by its rapid reaction with oxyhemoglobin to form methemoglobin and nitrate, monitored spectrophotometrically at 405 nm, offering a sensitive means to quantify cumulative NO• release from cells or tissues over time.[65] Similarly, NOS inhibitors like Nᴳ-nitro-L-arginine methyl ester (L-NAME) enable indirect assessment by blocking NO• production and observing reduced downstream effects, such as changes in cGMP levels, in cellular or organ preparations.[66]In environmental monitoring, denuder tubes facilitate the capture of gaseous NO-related species in ambient air by diffusive deposition onto coated inner surfaces, separating them from particulate matter for subsequent analysis. These annular or cylindrical tubes, often coated with alkaline substances like sodium carbonate, efficiently collect gaseous nitric acid (HNO₃) derived from NO oxidation, with collection efficiencies exceeding 95% at flow rates of 1-5 L/min, and are integrated into multi-stage samplers for comprehensive NOxspeciation in polluted atmospheres.[67]Basic chemiluminescence methods, adapted for field use, complement these assays by providing rapid on-site quantification of gaseous NO in environmental samples.[68]
Biological roles
Signaling and physiological functions
Nitric oxide (NO) is produced endogenously in mammalian cells primarily through the action of three isoforms of nitric oxide synthase (NOS) enzymes: neuronal NOS (nNOS), inducible NOS (iNOS), and endothelial NOS (eNOS). These enzymes catalyze the conversion of L-arginine to NO and L-citrulline in the presence of molecular oxygen (O₂) and the cofactor NADPH, with the overall reaction represented as:\text{L-arginine} + \text{O}_2 + \text{NADPH} \rightarrow \text{L-citrulline} + \text{NO} + \text{NADP}^+ + \text{H}_2\text{O}[69][70] nNOS, predominantly expressed in neurons, is calcium-calmodulin dependent and activated by neuronal depolarization, generating low levels of NO for localized signaling.[69] iNOS, found in immune cells such as macrophages, is calcium-independent and transcriptionally induced by proinflammatory stimuli like cytokines or lipopolysaccharide, producing sustained high-output NO fluxes.[70] eNOS, located in endothelial cells, is also calcium-dependent but tightly regulated by phosphorylation and shear stress, yielding pulsatile NO for vascular homeostasis.[69]A primary physiological function of NO is vasodilation, particularly mediated by eNOS-derived NO in the cardiovascular system. NO diffuses from endothelial cells into adjacent vascular smooth muscle cells, where it binds to the heme iron in soluble guanylate cyclase (sGC), activating the enzyme to catalyze the conversion of GTP to cyclic GMP (cGMP).[69] Elevated cGMP levels then activate protein kinase G, leading to dephosphorylation of myosin light chains and relaxation of smooth muscle, thereby reducing vascular tone and promoting blood flow.[70] This mechanism is essential for regulating blood pressure and ensuring adequate perfusion to organs, with disruptions in eNOS activity linked to endothelial dysfunction.[69]In the central nervous system, NO serves as a non-traditional neurotransmitter, primarily through nNOS activity in neurons. Unlike classical neurotransmitters stored in vesicles, NO is freely diffusible and acts in a paracrine manner, enabling long-range signaling across synaptic gaps and even between cells.[70] nNOS is activated by calcium influx via NMDA receptors during synaptic activity, and the resulting NO modulates postsynaptic processes such as long-term potentiation (LTP), a cellular correlate of learning and memory, by enhancing cGMP signaling in target neurons.[69] This diffusible nature allows NO to coordinate network-level activity in brain regions like the hippocampus and cerebellum.[70]NO also plays a critical role in the immune response, where iNOS in activated macrophages generates micromolar concentrations of NO to exert cytotoxic effects against pathogens. Upon stimulation by interferon-γ or bacterial products, iNOS expression is upregulated, leading to NO production that reacts with superoxide to form peroxynitrite, a potent oxidant that damages microbial proteins, lipids, and DNA.[69] This antimicrobial activity is exemplified in the defense against intracellular bacteria like Mycobacterium tuberculosis and Listeria monocytogenes, where iNOS-derived NO inhibits pathogen replication and promotes host survival.[70] Additionally, NO modulates inflammation by suppressing T-cell proliferation and cytokine release, fine-tuning the immune balance.[69]
Medical applications
Nitric oxide (NO) is utilized in medical practice primarily through inhalation therapy for treating persistent pulmonary hypertension of the newborn (PPHN). Inhaled NO acts as a selective pulmonary vasodilator, improving oxygenation by reducing pulmonary vascular resistance without significantly affecting systemic blood pressure. The U.S. Food and Drug Administration (FDA) has approved inhaled NO at a dose of 20 ppm for up to 14 days in term and near-term neonates with PPHN, with typical starting doses ranging from 5 to 20 ppm and potential escalation to 80 ppm based on response. Clinical trials have demonstrated that this therapy decreases the need for extracorporeal membrane oxygenation and enhances short-term oxygenation in affected infants.[71][72][73]NO donors, such as nitroglycerin, are widely employed in cardiovascular medicine, particularly for angina pectoris. Nitroglycerin is metabolized in vascular smooth muscle cells to release NO, which activates guanylate cyclase to increase cyclic guanosine monophosphate (cGMP) levels, leading to vasodilation and relief of myocardial ischemia. Administered sublingually, transdermally, or intravenously, nitroglycerin reduces preload and afterload, improving coronary blood flow and alleviating chest pain during acute episodes. This mechanism has established nitroglycerin as a cornerstone therapy for stable and unstable angina since the late 19th century.[74][75][76]During the COVID-19 pandemic, inhaled NO was investigated in clinical trials for acute respiratory distress syndrome (ARDS) associated with severe SARS-CoV-2infection. High-dose inhaled NO, up to 80 ppm, was tested to improve ventilation-perfusion matching and oxygenation in mechanically ventilated patients. A phase II multicenter trial showed that such dosing over 48 hours enhanced systemic oxygenation in critically ill individuals with hypoxemic respiratory failure, though it did not significantly reduce mortality. Some studies reported reduced ventilator dependence and shorter durations of mechanical ventilation in treated cohorts, particularly when initiated early, highlighting NO's potential adjunctive role in viral-induced lung injury.[77][78][79]Topical NO-releasing formulations have shown promise in promoting wound healing by modulating inflammation, enhancing angiogenesis, and providing antimicrobial effects. NO donors in creams or gels accelerate reepithelialization and collagen deposition in chronic wounds, such as diabetic ulcers and pressure sores, with clinical evidence indicating faster closure rates compared to standard care. For erectile dysfunction, therapies targeting the NO-cGMP pathway include phosphodiesterase-5 (PDE5) inhibitors like sildenafil, which prolong cGMP-mediated smooth muscle relaxation in the corpus cavernosum following NO release from endothelial cells during sexual stimulation. Topical NO gels, such as those delivering NO via microparticles, have demonstrated synergistic effects with oral PDE5 inhibitors in preclinical models, improving erectile responses without systemic side effects.[80][81][82][83]
Environmental occurrence and effects
Atmospheric chemistry
In the troposphere, nitric oxide (NO) is rapidly interconverted with nitrogen dioxide (NO₂) as part of the NOx (NOₓ = NO + NO₂) family, influencing ozone (O₃) levels through photochemical cycles. The primary reactions establishing this interconversion are:\text{NO} + \text{O}_3 \to \text{NO}_2 + \text{O}_2\text{NO}_2 + h\nu \to \text{NO} + \text{O}\text{O} + \text{O}_2 + \text{M} \to \text{O}_3 + \text{M}This sequence forms a null cycle with no net O₃ change in the absence of other radicals, but in polluted regions with volatile organic compounds (VOCs), hydroperoxyl (HO₂) and organic peroxy (RO₂) radicals convert NO to NO₂ while propagating O₃ formation, making NOx a key catalyst for tropospheric O₃ production.[84]In the stratosphere, NO contributes to O₃ destruction via a catalytic cycle involving atomic oxygen (O):\text{NO} + \text{O}_3 \to \text{NO}_2 + \text{O}_2\text{NO}_2 + \text{O} \to \text{NO} + \text{O}_2Net: \text{O}_3 + \text{O} \to 2\text{O}_2. This NOx cycle dominates O₃ loss in the mid-stratosphere, where NOx concentrations are elevated due to transport from the troposphere and in situ production from nitrous oxide (N₂O) oxidation.[85]The photostationary state (PSS) governs the daytime equilibrium among NO, NO₂, and O₃ in the troposphere, expressed as:k [\text{NO}][\text{O}_3] = J_{\text{NO}_2} [\text{NO}_2]where k is the rate constant for the NO + O₃ reaction (~9.9 × 10⁻¹⁴ cm³ molecule⁻¹ s⁻¹ at 298 K) and J_{\text{NO}_2} is the NO₂ photolysis rate (typically 0.01–0.02 s⁻¹ under solar conditions). This relation, [NO][O₃]/[NO₂] = J_{\text{NO}_2}/k, is fundamental for modeling urban smog formation and NOx-O₃ interactions in polluted air.[86]Global NOx sources total approximately 50–60 Tg N yr⁻¹, with anthropogenic emissions from fossil fuel combustion, industry, and agriculture contributing ~41 Tg N yr⁻¹ in recent years (e.g., 41.2 Tg N yr⁻¹ in 2023), while natural sources add ~10–20 Tg N yr⁻¹, primarily from soil microbes (~9.5 Tg N yr⁻¹) and lightning (~5 Tg N yr⁻¹). Sinks include oxidation to nitric acid (HNO₃) via OH radicals and wet/dry deposition, with NOx exhibiting a tropospheric lifetime of ~1 day (ranging 0.5–2 days depending on latitude and season).[87][88][39][89]In the 2020s, research emphasizes NOx's role in secondary organic aerosol (SOA) formation, particularly by altering the chemistry of biogenic VOC oxidation; under high-NOx conditions, it enhances nitroaromatic compounds and suppresses low-volatility products, contributing ~10–20% to global SOA burdens according to IPCC AR6 assessments.[90]
Pollution and ecosystem impacts
Nitric oxide (NO), as a component of nitrogen oxides (NOx), plays a significant role in the formation of acid rain through its atmospheric oxidation to nitrogen dioxide (NO₂), which subsequently reacts with water to produce nitric acid (HNO₃). The key reaction is 3NO₂ + H₂O → 2HNO₃ + NO, where nitric acid serves as a primary precursor to acidic precipitation.[91][92] This process contributes to the deposition of acidic compounds onto ecosystems, exacerbating soil and water acidification.[93]In aquatic systems, the wet deposition of nitrate ions (NO₃⁻) from NOx-derived nitric acid acts as a fertilizer, promoting eutrophication. Excess nitrogen inputs stimulate excessive algal growth, leading to blooms that deplete oxygen levels and create hypoxic zones harmful to aquatic life.[94][95] For instance, atmospheric NOx deposition has been identified as a major contributor to eutrophication in coastal waters, such as those in the Yellow Sea, where it intensifies nutrient overload.[94]On land, NOx emissions contribute to soil acidification, accounting for approximately 20-30% of acid rain acidity in regions like Europe and North America. This acidification mobilizes toxic aluminum in soils, inhibiting root growth and nutrient uptake in trees, which has been linked to forest dieback in temperate ecosystems.[96][97] Studies on nitrogen deposition effects highlight how chronic exposure weakens forest health, particularly in areas with high NOx emissions from industrial and vehicular sources.[98][99]Nitrogen overload from NOx deposition also drives biodiversity loss in sensitive terrestrial ecosystems, such as grasslands and heathlands. Elevated nitrogen levels alter soil chemistry, favoring nitrophilous (nitrogen-loving) plant species over native flora, which reduces species diversity and disrupts community structures.[100] In Europe, for example, critical nitrogen loads are often exceeded in these habitats, leading to long-term shifts in vegetation composition and ecosystem function.[95][101]
Climate and ozone influences
Nitric oxide (NO), primarily emitted from high-altitude aircraft exhaust, plays a significant role in stratospheric ozone depletion through catalytic cycles involving nitrogen oxides (NOx). In the stratosphere, NO reacts with ozone (O₃) to form nitrogen dioxide (NO₂) and molecular oxygen (O₂), while NO₂ subsequently reacts with atomic oxygen (O) to regenerate NO, resulting in the net destruction of ozone without net consumption of NOx. This catalytic cycle is represented as:\text{NO} + \text{O}_3 \rightarrow \text{NO}_2 + \text{O}_2\text{NO}_2 + \text{O} \rightarrow \text{NO} + \text{O}_2Net reaction: \text{O} + \text{O}_3 \rightarrow 2\text{O}_2. The process efficiently destroys odd-oxygen species (O and O₃) in the middle stratosphere (25–35 km altitude), where aircraft cruise emissions directly inject NOx into this sensitive region.[102]As an indirect greenhouse gas, NOx from anthropogenic sources contributes to climate forcing by enhancing tropospheric ozone concentrations, which act as a potent warming agent through radiative absorption. Additionally, NOx influences the atmospheric lifetime of methane (CH₄), a major greenhouse gas, by increasing hydroxyl radical (OH) concentrations that accelerate methane oxidation, thereby providing a counteracting cooling effect; however, the net radiative forcing from aviation NOx is positive due to the dominance of ozone production. This dual role underscores NOx's complex interplay in atmospheric chemistry, where stratospheric injections from aircraft can amplify both ozone loss and indirect warming pathways.High-altitude aviation emissions of NOx are estimated to contribute approximately 3–5% to total anthropogenic stratosphericozone loss, based on Federal Aviation Administration (FAA) atmospheric models that account for current subsonic fleets and projected growth. These models highlight that while the global impact remains modest compared to historical chlorofluorocarbon-driven depletion, localized depletions near flight corridors can exceed 10% in the upper stratosphere, influencing radiative balance and ultraviolet radiation penetration. Recent assessments confirm that aviationNOx perturbations alter the ozone column by 0.3–1% globally, with implications for surface climate through altered stratospheric dynamics.[103]In the 2020s, updated climate models incorporating NOx feedbacks from rising nitrous oxide (N₂O) emissions— the primary stratospheric source of NOx—indicate an amplification of global warming by 0.1–0.2°C through combined direct N₂O forcing and indirect ozone depletion effects. These feedbacks arise as increased N₂O photolysis produces more NOx, exacerbating catalytic ozone loss (cooling) while the potent greenhouse effect of N₂O dominates, leading to net warming; projections suggest this could account for up to 7% additional ozone reduction by mid-century under high-emission scenarios.[104] Such model refinements emphasize the need for integrated mitigation of NOx precursors to curb both ozone and climate risks.[102]
Safety and hazards
Occupational health effects
Occupational exposure to nitric oxide (NO) primarily occurs through inhalation in industrial settings, such as welding, metal fabrication, and agricultural operations, where it is generated as a byproduct of high-temperature processes or chemical reactions.[105] At low concentrations, NO is relatively non-irritating compared to its oxidation product nitrogen dioxide (NO₂), but it contributes to the toxicity of nitrogen oxides (NOx) mixtures.[106]Acute exposure to high levels of NO, typically above 100 ppm, can lead to methemoglobinemia by oxidizing hemoglobin to methemoglobin, which impairs oxygen transport in the blood and results in symptoms such as headache, fatigue, dizziness, cyanosis, and potentially respiratory failure or death.[105] Additional effects include pulmonary edema, pneumonia, and severe respiratory distress, with the immediately dangerous to life or health (IDLH) concentration established at 100 ppm.[107] These effects are exacerbated when NO rapidly oxidizes to NO₂ in the presence of oxygen, amplifying irritation to the respiratory tract.[106]Chronic exposure to lower levels of NO, often in the range of 1-5 ppm as part of NOx mixtures, is associated with respiratory irritation, including cough, shortness of breath, and bronchitis-like symptoms with increased phlegm production.[8] Prolonged inhalation may contribute to reduced lungfunction and chronic lung damage, particularly in occupational environments where cumulative exposure occurs over time.[108] To mitigate these risks, the Occupational Safety and Health Administration (OSHA) sets a permissible exposure limit (PEL) of 25 ppm as an 8-hour time-weighted average (TWA), with the National Institute for Occupational Safety and Health (NIOSH) recommending the same as a recommended exposure limit (REL).[108]Certain occupational groups face heightened risks due to frequent exposure. Welders, particularly those using arc or acetylene processes, inhale NO generated from atmospheric nitrogen reactions, leading to potential methemoglobinemia and respiratory issues.[109] Farmers are vulnerable during silo filling, where NO produced by fermenting silage oxidizes to toxic NO₂, causing "silo-filler's disease" with acute pulmonary edema and long-term lung impairment.[110] Regarding carcinogenicity, nitrogen oxides (NOx) mixtures have been evaluated, but the International Agency for Research on Cancer (IARC) has not classified nitric oxide with respect to its carcinogenicity to humans.[106]
Explosion and reactivity risks
Nitric oxide exhibits significant reactivity that can lead to explosive hazards, particularly in the presence of oxygen and combustible materials. It reacts rapidly with atmospheric oxygen to form nitrogen dioxide (NO₂), which exists in equilibrium with its dimer, dinitrogen tetroxide (N₂O₄). At high pressures, N₂O₄ acts as a powerful oxidizer and can decompose explosively, posing risks in confined systems such as during industrial handling or storage failures.[111][112]Mixtures of nitric oxide with hydrogen or hydrocarbons form highly reactive combinations capable of detonation over broad concentration ranges. For instance, in hydrogen-oxygen systems, the addition of NO broadens the explosion limits, enabling ignition and propagation at lower temperatures initially before shifting to higher thresholds as NO concentration increases, with detonation possible across non-monotonic Z-shaped response curves. Similar risks apply to hydrocarbon fuels, where NO sensitizes mixtures for rapid combustion or deflagration-to-detonation transitions.[113][114]Thermal decomposition of nitric oxide, following the reaction$2\text{NO} \rightarrow \text{N}_2 + \text{O}_2occurs readily above 1000°C, releasing nitrogen and oxygen gases in an endothermic process that can accelerate under shock or confinement, contributing to pressure buildup and potential explosions in overheated systems.[115]To mitigate these risks, nitric oxide is stored as a compressed gas in specialized cylinders designed for high-pressure containment, placed in cool, well-ventilated, and fireproof areas away from heat sources, combustibles, and incompatibles like ozone or halogens; rapid heating of containers can cause violent rupture.[116][8]
Regulatory standards
Regulatory standards for nitric oxide (NO) and related nitrogen oxides (NOx) are established by various international, regional, and national bodies to protect worker health and ambient air quality. In the United States, the Occupational Safety and Health Administration (OSHA) sets a permissible exposure limit (PEL) of 25 parts per million (ppm) as an 8-hour time-weighted average (TWA) for NO in workplace air.[108] The National Institute for Occupational Safety and Health (NIOSH) recommends a recommended exposure limit (REL) of 25 ppm TWA.[107]For ambient air quality, the U.S. Environmental Protection Agency (EPA) regulates nitrogen dioxide (NO₂) under the National Ambient Air Quality Standards (NAAQS) as a proxy for NOx exposure, including NO. The primary NAAQS for NO₂ is 100 parts per billion (ppb) over a 1-hour averaging period, based on the three-year average of the 98th percentile of daily maximum concentrations.[117] An annual standard of 53 ppb (100 µg/m³) also applies.[118]In the European Union, the Ambient Air Quality Directive (2008/50/EC, recast as Directive (EU) 2024/2881) establishes limit values for NO₂, which encompasses NOx emissions from sources producing NO. The current annual limit value is 40 µg/m³, with the revised directive halving it to 20 µg/m³ by 2030 to align more closely with health-based recommendations and integrate climate policy considerations, such as reducing emissions contributing to both air pollution and global warming.[119] The hourly limit value is 200 µg/m³, not to be exceeded more than 18 times per year, with the revision tightening the allowed exceedances to 3 times per calendar year from 1 January 2030.[120]Internationally, the World Health Organization (WHO) provides global air quality guidelines for NO₂ as a marker for traffic-related NOx, including NO. The 2021 guidelines recommend an annual mean concentration not exceeding 10 µg/m³ and a 1-hour mean of 200 µg/m³ to minimize health risks.[121] These guidelines serve as a benchmark for countries to develop or update their standards, emphasizing reductions in NOx emissions to protect vulnerable populations.[122]