Fact-checked by Grok 2 weeks ago

Nitrosonium

The nitrosonium cation (NO⁺) is a linear diatomic featuring a nitrogen-oxygen with a formal positive charge, exhibiting a of 3 and a short N-O of approximately 1.06 . This electrophilic species acts as a strong oxidant with a standard enabling it to oxidize substrates with potentials below about 1.7 V, and it serves as a key in , diazotization, and various processes in both and biological systems. Nitrosonium salts, such as the colorless crystalline nitrosonium tetrafluoroborate (NOBF₄), are commonly prepared by reacting with or through oxidation of , providing stable sources of NO⁺ for laboratory use. In , NO⁺ facilitates electrophilic additions to alkenes and aromatics, forming charge-transfer complexes, and is employed in the synthesis of diazonium salts from arylamines, heterocyclic compounds, and via catalytic and oxidation pathways. Biochemically, NO⁺ arises from the of (NO) in dinitrosyl iron complexes (DNICs) with ligands, enabling S-nitrosylation of residues in proteins to regulate signaling, , and cellular balance, though it can also contribute to and at elevated levels. Its short aqueous lifetime ( ~10⁻¹⁰ s at neutral ) underscores its role as a transient yet pivotal species in , such as stratospheric reactions contributing to .

Overview

Definition and Nomenclature

The nitrosonium ion is a diatomic cation with the NO⁺, formed by the removal of one from the neutral molecule (NO). This results in a species with a formal positive charge on the oxygen atom, exhibiting a character and a total of 10 valence electrons. The ion is isoelectronic with molecular (N₂), (CO), and the anion (CN⁻), sharing similar electronic configurations that contribute to its linear geometry and reactivity. In , the nitrosonium is commonly referred to by its , nitrosonium, which derives from the "" prefix associated with derivatives of (HNO₂). The is oxidonitrogen(1+). Other names include nitrilooxonium. Due to its high reactivity as a strong , the nitrosonium is rarely encountered in isolation and is instead stabilized in salts with weakly coordinating anions. Common counterions include tetrafluoroborate (BF₄⁻), (ClO₄⁻), and (PF₆⁻), which provide and minimal nucleophilic interaction to enable the preparation of crystalline, air-stable solids like nitrosonium tetrafluoroborate (NOBF₄). These anions play a crucial role in preventing decomposition or unwanted reactions, facilitating the ion's use in synthetic applications.

Physical Properties

Nitrosonium salts, such as nitrosonium tetrafluoroborate (NOBF₄) and nitrosonium (NOClO₄), are typically white or colorless crystalline solids. These compounds exhibit a around 2.2 g/cm³, with NOBF₄ specifically having a density of 2.185 g/cm³. These salts demonstrate high solubility in polar aprotic solvents, including and , facilitating their use in non-aqueous media. They react with water, undergoing . Regarding thermal stability, NOBF₄ sublimes at 200–250 °C under reduced pressure (0.01 mmHg) and decomposes before melting. Nitrosonium salts are hygroscopic and require storage under inert atmospheres, such as , in tightly sealed containers at low temperatures (2–8 °C) to prevent moisture absorption.

History

Discovery

The (NO⁺) was first proposed as a distinct by Arthur Rudolf Hantzsch in 1921, who suggested its existence in solutions of (NOCl), interpreting the compound's behavior as that of an ionic pair consisting of NO⁺ and Cl⁻ rather than a covalent . This proposal marked an early recognition of NO⁺ as a reactive cation capable of forming salts, based on and studies of NOCl in polar solvents, challenging prevailing views of nitrosyl compounds as neutral entities. Hantzsch's work laid foundational groundwork for understanding NO⁺ in non-aqueous media, highlighting its potential role in processes. A key experimental observation supporting the presence of NO⁺ came in 1923 from Frederick Daniel Chattaway and George Hoyle, who reported its formation in acidic solutions of (HNO₂). Their studies involved the reaction of with quaternary ammonium salts in strongly acidic conditions, yielding products consistent with the intervention of NO⁺ as an electrophilic intermediate, evidenced by the isolation of nitrosyl derivatives and conductance measurements indicating ionic dissociation. This experiment provided indirect evidence for NO⁺ in aqueous acidic environments, bridging theoretical proposals with practical observations. In the early , NO⁺ gained recognition as an implicated intermediate in diazotization reactions, where primary aromatic react with under acidic conditions to form diazonium salts. Pioneering kinetic and mechanistic studies during this period, building on Griess's original diazotization work from , pointed to NO⁺ as the active nitrosating agent responsible for transferring the group to the amine nitrogen, explaining reaction rates and product distributions in acidic media. Spectroscopic confirmation of the nitrosonium ion structure arrived in the through infrared studies of stable nitrosyl salts, such as NOBF₄ and NOClO₄. In 1953, John J. Turner reported characteristic IR absorption bands for NO⁺ between 2150 and 2400 cm⁻¹ in these compounds, attributing the strong, sharp peak near 2300 cm⁻¹ to the N≡O stretching vibration of the linear, triply bonded cation, providing definitive structural evidence and distinguishing it from neutral NO species. These findings solidified NO⁺'s identity and spurred further investigations into its coordination chemistry and reactivity.

Development of Stable Salts

The development of stable nitrosonium salts accelerated in the mid-20th century, enabling safer and more reliable access to the NO⁺ ion for synthetic applications. Earlier, unstable salts like (NOSO₄H) were known since the mid-19th century for use in diazotization reactions. Although earlier attempts to isolate nitrosonium compounds existed, the first truly stable salt, nitrosyl perchlorate (NOClO₄), was synthesized in the mid-1950s through the reaction of with , yielding a hygroscopic white solid that represented a milestone in the isolation of isolable NO⁺ species. This salt's preparation highlighted the challenges of handling perchlorates, which are prone to explosive decomposition, but it laid the foundation for subsequent advancements. To address the safety issues of NOClO₄, researchers introduced nitrosonium tetrafluoroborate (NOBF₄) in 1963 by J. P. Oliver and E. Griswold, obtained by treating nitrosyl chloride with boron trifluoride in an inert solvent. This colorless, crystalline salt offered superior stability and reduced hygroscopicity, making it a preferred reagent for laboratory use in nitrosation and mild oxidation reactions due to its easier handling and lower risk of detonation. In the 1970s, further progress focused on salts tailored for specialized applications, such as nitrosonium hydrogen sulfate (NOSO₄H) and nitrosonium hexafluorophosphate (NOPF₆), developed for electrochemical studies and processes. These compounds, prepared by metathesis reactions involving NO⁺ precursors and the corresponding anions, exhibited enhanced solubility in organic solvents and thermal stability, facilitating their use in non-aqueous electrochemistry and as one-electron oxidants. A key milestone in the 1980s was the expanded role of nitrosonium salts in , exemplified by G. Connelly's contributions to nitrosyl ligand transfer reactions. Connelly demonstrated that salts like NOBF₄ could selectively introduce NO ligands to centers, enabling the synthesis of nitrosyl complexes and the investigation of their electronic properties and reactivity, which advanced understanding of metal-nitrosyl bonding. Recent refinements have emphasized improving salt purity and stability for catalytic roles, often involving optimized techniques, which have enhanced the salts' performance in precise synthetic transformations, as verified by methods.

Structure and

Bonding and

The , denoted as \ce{NO^{+}}, features a with a between the and oxygen atoms, represented as \ce{N#O^{+}}, where the positive charge resides on the oxygen atom. This assignment yields formal charges of 0 on nitrogen (with five electrons: one and half of the six bonding electrons in the triple bond) and +1 on oxygen (with six valence electrons: one lone pair and half of the six bonding electrons). resonance forms include a double-bonded (\ce{^{+}N=O}), placing the positive charge on nitrogen, but the triple-bonded structure predominates due to octet satisfaction on both atoms. Molecular orbital theory describes the \ce{NO^{+}} bonding as a triple bond with an overall bond order of 3, comprising one \sigma bond from end-on orbital overlap and two \pi bonds from sideways p-orbital interactions, with all electrons paired in bonding and non-bonding orbitals. This high bond order results in a short N-O bond length of 1.06 Å, compared to 1.15 Å in neutral NO, reflecting increased electron density in the bonding region upon removal of the antibonding electron. The of \ce{NO^{+}} arises from hybridization on both and oxygen atoms, forming two orbitals that align collinearly (180° bond angle) for \sigma bonding, with unhybridized p orbitals accommodating the \pi bonds. This configuration mirrors the diatonic linearity of isoelectronic species like \ce{N2} and \ce{CO}, where the imparts comparable strengths, with dissociation energies around 1070 kJ/mol for \ce{NO^{+}} akin to those in \ce{N2} (945 kJ/mol) and \ce{CO} (1072 kJ/mol).

Spectroscopic Characteristics

The nitrosonium ion (NO⁺) exhibits a characteristic strong and sharp band due to the N≡O⁺ stretching vibration, observed in the range of 2150–2400 cm⁻¹ across various salts. This peak arises from the high and linear geometry of the ion, with specific frequencies varying slightly depending on the ; for instance, in nitrosonium tetrafluoroborate (NOBF₄), the band appears at approximately 2340 cm⁻¹. Raman spectroscopy similarly detects the NO⁺ stretching mode in solid salts, often at frequencies close to those in IR spectra, confirming the vibrational signature. In NOBF₄, the Raman-active stretch is observed at 2342 cm⁻¹, while in crown ether complexes like [18-crown-6·NOBF₄]⁺, it shifts to 2274 cm⁻¹ due to weak interactions with the ligand. In the ultraviolet-visible region, NO⁺ displays absorption near 140–150 nm, corresponding to the allowed π→π* electronic transition from the ground X¹Σ⁺ state to the excited A¹Π state, as determined from gas-phase spectroscopic studies. Mass spectrometry identifies the nitrosonium ion through its base peak at m/z 30, representing the intact NO⁺ species in ionization processes.

Synthesis

Laboratory Methods

The nitrosonium ion (NO⁺) can be generated in laboratory settings through the protonation of in strong acids, serving as an electrophilic species for subsequent reactions such as and diazotization. The reaction proceeds as follows: \text{HNO}_2 + \text{H}^+ \rightarrow \text{NO}^+ + \text{H}_2\text{O} This method is typically conducted using sulfuric acid (H₂SO₄) to provide the acidic medium, with nitrous acid formed in situ from sodium nitrite and the acid. Nitrosonium salts are commonly prepared by the oxidation of nitric oxide (NO) with halogens to form nitrosyl halides, followed by metathesis to yield the desired nitrosonium salt. For example, chlorine gas oxidizes NO to nitrosyl chloride: $2\text{NO} + \text{Cl}_2 \rightarrow 2\text{NOCl} The nitrosyl chloride is then treated with a silver salt for anion exchange, such as silver tetrafluoroborate to produce nitrosonium tetrafluoroborate: \text{NOCl} + \text{AgBF}_4 \rightarrow \text{NOBF}_4 + \text{AgCl} This two-step process allows for the isolation of stable nitrosonium salts under anhydrous conditions. A direct one-pot variant for NOBF₄ combines the oxidation and metathesis steps: \text{NO} + \frac{1}{2}\text{Cl}_2 + \text{AgBF}_4 \rightarrow \text{NOBF}_4 + \text{AgCl} This method is particularly useful for small-scale preparations where is not isolated. Another common laboratory method involves reacting (NO) with (BF₃), often in the presence of (HF) or other anhydrous media, to directly form NOBF₄. In addition, the nitrosonium ion can be generated via electrochemical oxidation of NO in non-aqueous solvents like , enabling controlled delivery for synthetic applications without isolating the salt. This approach involves anodic oxidation at a suitable , typically using electrodes.

Purification Techniques

Nitrosonium salts, such as nitrosonium tetrafluoroborate (NOBF₄), are highly hygroscopic and moisture-sensitive, necessitating purification under strictly and inert conditions to prevent . sublimation is a primary method for obtaining high-purity NOBF₄, typically performed at temperatures of 200–250 °C under reduced of approximately 0.01 mmHg (1.3 ), yielding colorless crystals free from contaminants like nitronium tetrafluoroborate. This technique exploits the compound's volatility without decomposition, achieving analytical-grade purity. Recrystallization from anhydrous solvents like or is commonly employed for further refinement, with the solution prepared under and cooled to -20 °C to induce , followed by in a to maintain inertness. These non-coordinating solvents minimize effects while dissolving impurities, resulting in enhanced crystalline quality. Metathesis reactions allow anion exchange to improve stability, for instance, converting less stable nitrosonium to the tetrafluoroborate by reaction with a silver or salt of the desired anion in a suitable , precipitating the more stable product. This approach is particularly useful for preparing salts with weakly coordinating anions that enhance handling and storage properties. Post-purification drying is essential to remove residual moisture or solvent; nitrosonium salts are often dried over phosphorus pentoxide (P₂O₅) in a desiccator under vacuum to ensure anhydrous conditions without direct contact to avoid side reactions. Vacuum drying at room temperature may also be applied for solvent removal after recrystallization. Purity is verified analytically using infrared (IR) spectroscopy, where the characteristic N-O stretching frequency at approximately 2300–2387 cm⁻¹ serves as a key indicator; a sharp, intense peak confirms the absence of hydrolysis products or impurities.

Reactivity

Hydrolysis and Aqueous Behavior

The nitrosonium ion undergoes rapid hydrolysis in aqueous solution via the reaction \ce{NO+ + H2O -> HONO + H+} with an experimental lifetime of approximately $3 \times 10^{-10} s, corresponding to a pseudo-first-order rate constant on the order of $3 \times 10^{9} s^{-1}. This process exhibits pH dependence, proceeding faster in neutral or less acidic conditions where the forward reaction is favored, while increased acidity shifts the system toward the reverse direction, extending the effective lifetime of NO^+. The instability of the nitrosonium ion extends to other protic s, where it similarly decomposes to form (HONO) through nucleophilic attack by the solvent oxygen. In acidic media, is reversible, establishing the \ce{NO+ + H2O <=> HNO2 + H+} which allows detectable concentrations of NO^+ under strongly acidic conditions ( < 1), as the proton concentration suppresses full decomposition. The lifetime and overall stability of nitrosonium salts in humid or aqueous environments are influenced by the counteranion; for instance, tetrafluoroborate (BF_4^-) salts exhibit greater resistance to compared to those with more nucleophilic anions, due to reduced ion-pairing and lower hygroscopicity, enabling their use in synthetic applications with minimal moisture exposure. The decay of nitrosonium in aqueous media has been studied computationally and experimentally, confirming the rapid conversion to .

Diazotization Reactions

Nitrosonium ions () serve as electrophilic agents in diazotization reactions, facilitating the conversion of primary aromatic amines to aryldiazonium salts, which are versatile intermediates in organic synthesis. The mechanism involves the initial electrophilic attack of on the nitrogen lone pair of the arylamine (), forming an N-nitrosoammonium intermediate, followed by protonation and dehydration to yield the diazonium ion (). This process can be represented by the equation: \text{ArNH}_2 + \text{NO}^+ + \text{H}^+ \rightarrow \text{ArN}_2^+ + \text{H}_2\text{O} A common reagent for generating in these reactions is (NOBF₄), which is employed under mild conditions, such as in acetonitrile at low temperatures (e.g., -40 °C), allowing for efficient diazotization without harsh acidic media. This approach offers advantages over the classical method, including cleaner reaction profiles, enhanced stability of the diazonium products, and avoidance of nitrite over-reduction or unwanted side products like azo compounds. A representative example is the diazotization of aniline to benzenediazonium tetrafluoroborate using NOBF₄, which provides a key precursor for azo dye synthesis through coupling with activated aromatic substrates like phenols or naphthols.

Oxidation Reactions

The nitrosonium ion (NO⁺) functions as a potent single-electron oxidant in redox processes, with a standard reduction potential of +1.28 V vs. SCE in acetonitrile. This enables selective electron removal from organic substrates, as exemplified by the half-reaction $2\text{NO}^+ + 2\text{e}^- \rightarrow 2\text{NO} which underscores its thermodynamic favorability for oxidizing species with lower reduction potentials. In ether cleavage, NO⁺ promotes C-O bond scission through electrophilic attack on the ether oxygen, generating activated intermediates that transform into carbonyl products upon hydrolysis. For instance, methyl ethers are converted to ketones using NOBF₄. The process highlights NO⁺'s role in activating otherwise stable C-O bonds. Recent applications include nitrosonium-catalyzed oxidative bromination of electron-rich arenes using bromide salts and O₂ (as of 2024). NO⁺ also facilitates the oxidation of oximes to nitro compounds, leveraging its electrophilic and oxidative properties to convert C=NOH functionalities into R-NO₂ groups, often in the presence of co-reagents like N-vinylpyrrolidone to stabilize intermediates. This transformation provides a route to nitroalkanes from readily available oximes, emphasizing conceptual control over regioselectivity in nitrogen-containing oxidations. Dehydrogenation reactions mediated by NO⁺ include the conversion of alcohols to aldehydes, typically in aerobic conditions with cupric ions. For example, primary alcohols (RCH₂OH) are oxidized to RCHO using O₂/Cu²⁺/NO⁺ systems, where NO⁺ acts as an electron mediator to facilitate hydride abstraction without over-oxidation. Similarly, NO⁺ supports dehydrogenation of amines to imines or further to nitriles by generating radical cations that lose hydrogen equivalents, as seen in secondary amine oxidations to C=N bonds. These processes exemplify NO⁺'s utility in mild, selective dehydrogenations for synthetic applications.

Nitrosylation of Organic Substrates

Nitrosonium ion (NO⁺) serves as a potent electrophile in the nitrosylation of organic substrates, particularly through electrophilic aromatic substitution (EAS) reactions with electron-rich arenes, leading to the formation of C-nitroso compounds. This process involves the direct addition of the NO group to the carbon framework of the aromatic ring, distinct from N-nitrosation pathways. The general reaction can be represented as: \text{ArH} + \text{NO}^+ \rightarrow \text{ArNO} + \text{H}^+ where ArH denotes an aromatic hydrocarbon. The mechanism proceeds via the attack of NO⁺ on the π-electron system of the arene, generating a Wheland intermediate (σ-complex) in the rate-determining step. This positively charged intermediate undergoes deprotonation to restore aromaticity, followed by tautomerization of the initial oxime-like structure to the thermodynamically stable nitroso form (ArNO). The deprotonation step exhibits reversibility, as evidenced by kinetic isotope effects, and is influenced by the electron-withdrawing nature of the NO group, which destabilizes the intermediate more than in analogous nitrations. Reactivity is highly selective for electron-rich arenes, with para-directing groups such as alkoxy or hydroxy substituents significantly enhancing the rate and directing substitution to the para position due to stabilization of the Wheland intermediate. For instance, undergoes nitrosylation predominantly at the para position to yield in high regioselectivity under mild conditions using NO⁺ sources like . Representative examples include the nitrosation of phenols, which readily form para-nitroso derivatives owing to the strong activating effect of the hydroxy group, and indoles, where substitution occurs preferentially at the C3 position of the pyrrole ring. These reactions highlight NO⁺'s utility in functionalizing biologically relevant heterocycles without requiring harsh conditions. Nitrosonium-initiated C-H activation enables synthesis of homo-diarylamines from arenes (as of 2024).

Formation of Nitrosyl Complexes

Nitrosonium ion (NO⁺) serves as an effective transfer reagent for introducing the nitrosyl ligand into metal complexes, facilitating the formation of metal-nitrosyl bonds through electrophilic attack on metal centers or ligand displacement. In a typical reaction, NO⁺ interacts with a metal-ligand complex (M-L) to yield a cationic nitrosyl species (M-NO⁺) and the displaced ligand (L), often in coordinating solvents like acetonitrile or methanol. This process is exemplified by the oxidation of chromium hexacarbonyl with nitrosonium-based oxidants, such as [NO]⁺[Al(OR_F)_4]⁻ (R_F = C(CF_3)_3), which generates [Cr(CO)_6]⁺ initially, followed by NO/CO ligand exchange to form the stable 18-electron [Cr(CO)_5NO]⁺ complex in near-quantitative yield under room-temperature conditions in dichloromethane. Similarly, complexes like (η^5-C_5H_5)Fe(CO)_2NO can act as sources of NO⁺ equivalents, enabling transfer to other metals via oxidative pathways that promote nitrosylation. The geometry of the nitrosyl ligand in these complexes varies between linear and bent configurations, influencing their electronic properties and reactivity. Linear M-N-O arrangements predominate when NO behaves as a two-electron donor (NO⁺), featuring a strong N≡O triple bond, while bent geometries arise when NO acts as a one-electron donor (NO^• or NO⁻), resulting in a weaker N=O double bond with greater π-backbonding from the metal. This distinction is captured by the Enemark-Feltham notation, {MNO}^n, where n represents the total electrons in the metal d orbitals plus the NO π* orbitals; linear forms typically exhibit n ≤ 6–7, whereas bent forms have n ≥ 8. Spectroscopic methods, particularly infrared (IR) spectroscopy, distinguish these ligand types by their characteristic N-O stretching frequencies (ν_NO). In metal-nitrosyl complexes with linear NO⁺, ν_NO shifts to higher energies in the 1700–1900 cm⁻¹ range due to reduced backbonding and increased N-O bond order, compared to free NO⁺ at approximately 2224 cm⁻¹; bent NO ligands, conversely, show lower frequencies around 1525–1690 cm⁻¹. Nitrosonium-mediated nitrosylation finds application in the preparation of polynuclear nitrosyl clusters, such as those in , where Fe-S frameworks incorporate multiple NO ligands to form species like [Fe_2S_2(NO)_4]^{2-} (red salt) and [Fe_4S_3(NO)_7]^- (black salt), classified under as {FeNO}^7 units with mixed linear and bent geometries. In catalysis, NO⁺ promotes ligand substitution in organometallic systems by oxidizing labile ligands (e.g., halides or CO), enabling selective replacement with NO or ancillary groups to generate active species for processes like carbonylation or hydrogenation.

Applications

Organic Synthesis

Nitrosonium salts, particularly nitrosonium tetrafluoroborate (NOBF₄), serve as efficient reagents for the diazotization of aromatic amines, enabling the formation of aryldiazonium salts that are key intermediates in . In this process, anilines react with NOBF₄ in organic solvents such as acetonitrile or dichloromethane at low temperatures (0–5 °C) to generate stable aryldiazonium tetrafluoroborates (ArN₂⁺ BF₄⁻) without the need for strong acids like HCl, avoiding side reactions such as hydrolysis. These diazonium salts then undergo copper-mediated substitution with halides (CuX, where X = Cl, Br, I) or cyanide (CuCN) to yield aryl halides (ArX) or aryl nitriles (ArCN), respectively, with yields often exceeding 80% for electron-rich and -neutral substrates. This approach provides a safer alternative to traditional nitrous acid-based diazotization, as NOBF₄ is stable and handles well, minimizing the risks associated with explosive diazonium intermediates. In nitrosation reactions, nitrosonium ions facilitate the introduction of nitroso groups into organic substrates, contributing to the synthesis of nitroso dyes and pharmaceutical intermediates such as furoxans. For nitroso dyes, NO⁺ reacts with activated aromatic systems like phenols or anilines to form C-nitroso compounds, which can be further elaborated into colored derivatives used in textile applications; for instance, nitrosation of p-cresol yields p-methylnitrosophenol, a precursor to azo-nitroso hybrids with enhanced lightfastness. In pharmaceutical contexts, NOBF₄ enables the regioselective synthesis of furoxans (1,2,5-oxadiazole 2-oxides) from styrenes in non-acidic media such as pyridine or dichloromethane at room temperature, proceeding via double nitrosation and cyclization to afford the heterocycles in moderate to good yields (50–75%) while tolerating acid-sensitive groups like esters. Furoxans act as nitric oxide (NO) donors in drugs targeting vasodilation and anti-inflammatory effects, highlighting nitrosonium's role in constructing bioactive motifs. As an oxidant, nitrosonium promotes dehydrogenative couplings and selective alcohol oxidations under mild conditions. In cross-dehydrogenative couplings, catalytic NO⁺ (generated from NOBF₄ or NaNO₂) facilitates metal-free C–S and C–N bond formation between electron-rich arenes (e.g., indoles, phenols) and thiols or amines under aerobic conditions at room temperature, using air as the terminal oxidant and producing water as the sole byproduct; for example, the coupling of indole with thiophenol yields the C3-thioether in 77% yield, demonstrating broad substrate scope (>60 examples) and high without prefunctionalization. For alcohol oxidations, nitrosonium mediates the conversion of primary alcohols to aldehydes using O₂ and Cu(II) ions, where NO⁺ acts as a transient one-electron oxidant (E° = 1.28 V vs. ) to form the carbonyl via sequential hydrogen abstractions, achieving quantitative yields for benzylic alcohols like to in at ambient temperature. These transformations underscore nitrosonium's versatility as a mild, selective oxidant compared to harsher agents like Cr(VI). Recent developments include nitrosonium-catalyzed oxidative bromination of arenes (as of 2024). Recent advancements in the have integrated nitrosonium into chemistry for safer diazotization, addressing scalability and hazard concerns in batch processes. For instance, NOBF₄-mediated diazotization of anilines in continuous- microreactors allows precise control of (seconds to minutes) and temperature, generating diazonium salts for immediate Sandmeyer-type substitutions, with reported yields up to 90% for aryl chlorides while minimizing accumulation of unstable intermediates. This method enhances by enabling small-scale handling of reactive and has been applied in the immobilization of functional groups on , where diazotized intermediates couple efficiently under conditions. Overall, nitrosonium's advantages in include high selectivity for electron-rich sites, compatibility with mild and non-aqueous conditions, and reduced environmental impact through catalytic turnover and benign byproducts, making it preferable to traditional acidic or metal-heavy protocols.

Coordination and Inorganic Chemistry

The nitrosonium ion (NO⁺) is a versatile reagent in coordination and , primarily functioning as a nitrosyl ligand source, one-electron oxidant, and halide abstractor in the synthesis and modification of complexes. Its linear and strong π-acceptor properties facilitate the formation of stable {M–NO}⁶ units, where the superscript denotes the total electron count contributed by the metal-nitrosyl fragment according to the Enemark-Feltham notation. NO⁺ salts, such as nitrosonium tetrafluoroborate (NOBF₄) or (NOPF₆), are commonly employed due to their in solvents and clean reactivity profiles. As an oxidant, NO⁺ promotes the conversion of low-valent transition metals to higher oxidation states by transferring its electron, often concomitant with nitrosyl ligation. For instance, treatment of neutral ((η⁵-C₅H₅)₂Fe) with NO⁺ yields the ferrocenium cation ((η⁵-C₅H₅)₂Fe⁺), demonstrating selective one-electron oxidation without incorporation in some cases. Similarly, the dinuclear thiolate-bridged [η⁵-C₅H₅Fe(CO)SCH₃]₂ is oxidized to [η⁵-C₅H₅Fe(CO)SCH₃]₂⁺, preserving the core structure while altering its properties. A representative example involves (II) ammine precursors, where NO⁺ oxidizes Ru²⁺ to Ru³⁺ while introducing the nitrosyl , affording [Ru(NH₃)₅NO]³⁺; this serves as a model for storage and release in biological systems. These processes are particularly useful for tuning the electronic properties of metal centers in catalytic applications. In the formation of polynitrosyl clusters, NO⁺ contributes to the assembly of multidentate nitrosyl frameworks, notably in dinitrosyl iron complexes (DNICs). Binuclear DNICs, such as [(GS⁻)₂Fe₂(NO)₄] (GS⁻ = glutathionyl), incorporate NO⁺ equivalents that enable S-nitrosation of thiols, mimicking enzymatic NO transfer. These clusters form via reaction of Fe²⁺ salts with nitrite sources that generate NO⁺ in situ, leading to {Fe(NO)₂}⁹ units where each NO ligand exhibits partial NO⁺ character due to antiferromagnetic coupling and charge delocalization; decomposition in acidic media releases 50% of the nitrosyls as NO⁺. Such polynitrosyl species provide insights into NO-mediated signaling in bioinorganic contexts. NO⁺-mediated halide abstraction is a key strategy in organometallic synthesis for generating cationic intermediates or substituting labile ligands. Reaction of NOPF₆ with halocarbonyl complexes promotes removal, often with concomitant solvent coordination. For example, η⁵-C₅H₅Fe(CO)₂I reacts with NO⁺ in to form [η⁵-C₅H₅Fe(CO)₂(CH₃CN)]PF₆, effectively replacing with a weakly bound ligand. Analogous abstraction occurs in and systems, such as (η⁵-C₅H₅)₂VCl₂ yielding [(η⁵-C₅H₅)₂V(CH₃CN)₂][PF₆]₂ and Mn(CO)₅Br giving [Mn(CO)₅(CH₃CN)]PF₆, facilitating further reactivity at the metal center. This method is preferred over traditional silver salts for its milder conditions and avoidance of insoluble precipitates. Ligand transfer involving NO⁺ is instrumental in constructing bioinorganic models, particularly heme-NO mimics. Synthetic iron porphyrin complexes, such as Fe(TPP) (TPP = ), react with NO⁺ sources to form nitrosyl adducts that mimic the {Fe–NO}⁷ configuration in ferrous heme-nitrosyls, enabling studies of NO binding and release. These model complexes demonstrate efficient NO transfer to acceptor hemes, analogous to biological shuttling in activation, where the thermodynamic of NO disrupts proximal histidine coordination. Such transfers highlight NO⁺'s role in simulating enzymatic NO delivery without direct organic substrate involvement.

Industrial and Analytical Uses

Nitrosonium ion serves as a key in the industrial diazotization of aromatic amines to produce diazonium salts, which are essential intermediates for synthesizing azo dyes and pigments on a large scale. These azo compounds, formed via subsequent reactions, account for a significant portion of synthetic colorants used in textiles, inks, and plastics, with global exceeding 700,000 tons annually. The process typically employs salts in acidic media to generate nitrosonium , ensuring efficient conversion under controlled conditions to meet high-volume demands. Commercial nitrosonium salts, such as nitrosonium tetrafluoroborate (NOBF₄), are widely utilized in the preparation of pharmaceutical intermediates through selective reactions. For instance, NOBF₄ enables the ipso-nitrosation of organotrifluoroborates, yielding stable derivatives that serve as building blocks for drug candidates, offering advantages in yield and purity over traditional methods. This application supports the synthesis of complex molecules in pipelines. In , nitrosonium salts function as precise reagents for the determination of primary aromatic amines via diazotization followed by , producing measurable colored products for spectrophotometric quantification. NOBF₄, in particular, provides clean generation of nitrosonium without excess acid, enhancing accuracy in trace-level analysis of environmental and biological samples. Additionally, gallium oxide-based electrochemical sensors offer high sensitivity and selectivity for detecting oxides in gas streams for industrial monitoring. Recent advancements post-2020 highlight nitrosonium's role in , particularly as a recyclable oxidant in sustainable processes like the nitro-oxidation of cellulosic materials under mild conditions (as of 2025), minimizing waste and enabling valorization. In industrial scale-up, NOBF₄ is favored over gaseous (NOCl) due to its stable solid form, which reduces handling hazards, risks, and exposure to toxic fumes, thereby improving safety and operational efficiency.

References

  1. [1]
    The Biologically Relevant Coordination Chemistry of Iron and Nitric ...
    Dec 13, 2021 · ... nitrosonium ion (NO+, ∼1.06 Å) and the nitroxyl anion (3NO–, ∼1.20 ... bond length was fixed at 1.15 Å during structure refinement).
  2. [2]
    [PDF] Nitrosonium cation in chemical and biochemical reactions
    Apr 27, 2016 · A nitrosonium cation is a reactive intermediate in numerous chemical and biochemical reactions. The NO+ cation is pertinent to the chemistry ...
  3. [3]
    Ion radicals. 37. Preparation and isolation of cation radical ...
    37. Preparation and isolation of cation radical tetrafluoroborates by the use of nitrosonium tetrafluoroborate.
  4. [4]
    How is Nitric Oxide (NO) Converted into Nitrosonium Cations (NO+) ...
    Oct 20, 2020 · In the acidic environment, both NO+–NO2− adduct components should, in the absence of thiols, convert into two HNO2 molecules via a hydrolysis ...Missing: applications | Show results with:applications
  5. [5]
    Relationships between nitric oxide, nitroxyl ion, nitrosonium cation ...
    Loss of the unpaired electron gives NO+, which has a bond order of 3 and is isoelectronic with N2 and CO. Nitric oxide is difficult to oxidise, with a ...
  6. [6]
    nitrosonium, n. meanings, etymology and more
    OED's earliest evidence for nitrosonium is from 1934, in Chemical Abstracts. nitrosonium is formed within English, by compounding. Etymons: nitroso- comb. form, ...Missing: ion | Show results with:ion
  7. [7]
    CHEBI:29120 - oxidonitrogen(1+) - EMBL-EBI
    Jul 27, 2012 · IUPAC Name. oxidonitrogen(1+). Synonyms, Sources. [NO]+, MolBase. nitric oxide(1+), ChEBI. nitrilooxonium, IUPAC. Nitrogen oxide cation, NIST ...
  8. [8]
    Reactive nitrogen species: Nitrosonium ions in organic synthesis
    Mar 1, 2019 · Nitrosonium ions can be generated using nitrosonium salts with different counterions (e.g. BF4−, PF6−, HSO4−, ClO4−), nitrous acid ...
  9. [9]
    NITROSONIUM TETRAFLUOROBORATE | 14635-75-7
    Chemical Name: NITROSONIUM TETRAFLUOROBORATE. Synonyms: NITROSYL ... Nitrilooxonium tetrafluoroborate azanylidyneoxidanium:tetrafluoroborate ...Missing: IUPAC | Show results with:IUPAC<|separator|>
  10. [10]
    Buy Nitrosonium perchlorate | 15605-28-4 - Smolecule
    Rating 4.0 (9) It appears as a hygroscopic white solid and is characterized as a salt formed from the nitrosonium cation NO + NO+ and the perchlorate anion ClO 4 − ClO4−​.
  11. [11]
  12. [12]
    Buy Nitrosonium tetrafluoroborate (EVT-314985) | 14635-75-7
    Initial investigations focused on stabilizing reactive cations through non-nucleophilic counterions, with George Andrew Olah's pioneering work on ...
  13. [13]
    Synthesis and Thermal Decomposition Studies of New Nitroso
    The alkali metal, silver, and barium salts decompose at higher temperatures (>200 °C), whereas the nitrogenous cationic salts decompose at lower temperatures, ...
  14. [14]
    [PDF] SAFETY DATA SHEET - Fisher Scientific
    Sep 26, 2009 · Do not get in eyes, on skin, or on clothing. Storage. Keep in a dry place. Keep container tightly closed. Keep under nitrogen. Keep refrigerated ...
  15. [15]
  16. [16]
    Formation, Structure, and Properties of Nitrosonium and Nitronium ...
    Complex reactions are found when N2O3 and N2O4 are used as substrates, resulting in mixed or impure products. An explanation involving the intermediate ...
  17. [17]
    Chemical Redox Agents for Organometallic Chemistry
    The nitrosonium ion is a strong oxidant {e.g., it oxidizes [N(C6H4Br-4)3] (E°' = 0.70 V) to [N(C6H4Br-4)3]+, section II.B.1.a)}. Reactions with both ...
  18. [18]
    Reactive nitrogen species: Nitrosonium ions in organic synthesis
    Jan 29, 2019 · Nitrosonium ions are versatile and mild oxidants, which were successfully employed in organic transformations. The applications cover different fields.
  19. [19]
    NO+ lewis structure, molecular geometry, bond angle, hybridization
    Jun 10, 2025 · NO+ has sp hybridization. The N≡ O bond length is equal to 106 pm in the nitrosonium ion. NO+ is a polar cation as the charged electron ...
  20. [20]
    Structures for NO+ (Nitrosonium) - Chemistry Stack Exchange
    Sep 19, 2015 · You could think of it as a 2.9-bond. Molecular orbital theory tells us that the actual bond order in NO+ is 3, but that's an entirely separate ...
  21. [21]
    List of experimental bond lengths for bond type rN=O - CCCBDB
    List of experimental bond lengths for bond type rN=O ; rN=O · ClNO, Nitrosyl chloride, 1.139 ; rN=O · N2O · Dinitrogen trioxide, 1.142, NO end.Missing: NO+ | Show results with:NO+
  22. [22]
    670. The Infrared Spectrum of the Nitrosonium Ion. - RSC Publishing
    THORLEY. Compounds containing the nitrosonium ion have an infrared absorption peak between 2150 and 2400 cm. -l. The changes in absorption frequency are ...
  23. [23]
    1 stoichiometric complexes of nitronium and nitrosonium ...
    The stretching vibration of the NO' ion usually occurs in the 2300-2350 cm-' frequency range, e.g. 2340 cm-' in NOBF, (9), 2326 cm-' in NOAuF6 (1 I), and 2298 ...Missing: spectroscopy | Show results with:spectroscopy
  24. [24]
    The valence states of NO+ - ScienceDirect.com
    Spectroscopic constants for the eight lowest electronic states of the NO+ ion are tabulated. These constants result from reanalysis of previously reported ...
  25. [25]
  26. [26]
    Protonation of nitrous acid and formation of the nitrosating agent NO+
    Protonation is predicted to occur on the hydroxylic oxygen to give (4) as the most stable structure. This shows an unusually long N–O(2) bond distance and is ...Missing: H2SO4 laboratory
  27. [27]
    Direct Nitrosation of Aromatic Hydrocarbons and Ethers with the ...
    Direct Nitrosation of Aromatic Hydrocarbons and Ethers with the Electrophilic Nitrosonium Cation ... Nitrosonium Ion Catalyzed Oxidative Bromination of Arenes.
  28. [28]
    Flow Electrochemistry for the N‐Nitrosation of Secondary Amines - Ali
    Mar 28, 2023 · Typically, a compound containing an electrophilic nitrosonium ion [(NO+)], that is, a nitrosating agent, is required to access N-nitrosamines.
  29. [29]
    [PDF] 16.5. Preparation and Reactions of Nitrosonium Tetrafluoroborate
    Nitrosonium tetrafluoroborate (1) is a white powder made from hydrogen fluoride, boron trifluoride and nitrogen dioxide.
  30. [30]
    Onium ions. 20. Ambident reactivity of the nitronium ion. Nitration vs ...
    Nitric acid in the presence of P2O5 supported on silica gel—a useful reagent ... Nitrosonium complexes of organic compounds. Structure and reactivity ...
  31. [31]
    Ab initio molecular dynamics simulation of aqueous solution of nitric ...
    Oct 1, 2015 · Nitrosonium cation has a shorter lifetime in aqueous solution than predicted experimentally. Nitrosonium cation in aqueous solution promptly ...Missing: instability | Show results with:instability
  32. [32]
    Intracluster reaction dynamics of NO+(H2O)n - AIP Publishing
    Sep 4, 2024 · At n = 4, HONO was formed by IR photon excitation. Eyet et al. measured the rate constants of H2O clustering to NO+(H2O)1– ...
  33. [33]
  34. [34]
  35. [35]
    Oxidation of alcohols to aldehydes with oxygen and cupric ion ...
    Oxidation of alcohols to aldehydes with oxygen and cupric ion, mediated by nitrosonium ion | Journal of the American Chemical Society.
  36. [36]
    Preparations of C-Nitroso Compounds | Chemical Reviews
    An early preparation 20 of dimeric 4-nitrosomenthon employed pentyl nitrite and HCl as the nitrosating agent, and such nitrosations have been extended using ...
  37. [37]
  38. [38]
  39. [39]
  40. [40]
  41. [41]
    [PDF] Nitrosonium Ions as Constituents of Dinitrosyl Iron Complexes with ...
    Nov 12, 2018 · Abstract. It has been established that the S-nitrosating activity of biologically active binuclear dinitrosyl iron complexes with ...
  42. [42]
    Classifications, properties, recent synthesis and applications of azo ...
    Jan 31, 2020 · The azo dyes and pigments are manufactured on an industrial scale by the same reaction sequence in two stages, diazotization and azo coupling.
  43. [43]
    PMC - NIH
    The nitrosation of organotrifluoroborates uses nitrosonium tetrafluoroborate (NOBF4) for ipso-nitrosation, converting them into nitroso products. This is the ...
  44. [44]
    [PDF] NITROGEN OXIDES GAS SENSOR UTILIZING MICROPOROUS ...
    Aug 16, 2023 · This sensor was interesting in that a new nitrosonium ion (NO+) conducting Ga11O17 electrolyte was utilized. The two electrodes were made of ...
  45. [45]
    Pressure effect on nitro-oxidation process for converting cellulosic ...
    The reaction between HNO3 and KNO2 can generate nitrosonium ions (NO+), which selectively oxidize the hydroxyl groups at the C6 carbon position of the beta-d- ...Pressure Effect On... · 2. Materials And Methods · 3. Results And Discussion