Nitrogen
Nitrogen is a chemical element with the symbol N and atomic number 7, belonging to group 15 (pnictogens) and period 2 of the periodic table.[1] It is a colorless, odorless, and tasteless nonmetal that occurs primarily as the diatomic gas N₂ at standard temperature and pressure, making up approximately 78.1% of Earth's atmosphere by volume and serving as the most abundant element in the air.[1][2] With an atomic mass of 14.007 u and electron configuration [He] 2s² 2p³, nitrogen exhibits five valence electrons, enabling it to form strong triple bonds in N₂ and adopt oxidation states ranging from -3 to +5 in compounds.[1][3] In terms of physical properties, nitrogen gas has a density of 0.001145 g/cm³ at 20°C, a melting point of -210.00°C, and a boiling point of -195.79°C, allowing it to be liquefied for cryogenic applications.[1][3] Chemically, N₂ is relatively inert due to the high bond energy of its triple bond (941 kJ/mol), which contributes to its role as an asphyxiant and diluent in respiration, though it is essential for life through incorporation into biological molecules.[2] Despite its atmospheric abundance, nitrogen is scarce in Earth's crust at about 1.9 × 10¹ mg/kg, primarily occurring in nitrates, ammonia, and organic forms.[1] Nitrogen plays a critical role in biology as a key component of amino acids, proteins, nucleic acids (such as DNA and RNA), and other biomolecules, cycling through ecosystems via the nitrogen cycle involving fixation by bacteria and plants.[1][2] Industrially, it is vital for the Haber-Bosch process to produce ammonia for fertilizers, supporting global food production, and is used in nitric acid synthesis for explosives and chemicals, as well as an inert atmosphere in metallurgy, food packaging, and semiconductor manufacturing.[1][2] Liquid nitrogen also finds applications in cryogenics, medical freezing, and food preservation due to its low temperature.[1]History
Discovery and isolation
Nitrogen was first isolated in 1772 by Scottish physician Daniel Rutherford during his doctoral research at the University of Edinburgh. Rutherford conducted experiments in which he confined a mouse in a sealed container of air until it suffocated, then removed the carbon dioxide (referred to as "fixed air") produced by respiration using an alkali solution, such as limewater. The remaining gas, which he termed "mephitic air" or "noxious air," extinguished flames and was lethal to small animals, leading him to conclude it was a distinct component of the atmosphere saturated with phlogiston, separate from oxygen and carbon dioxide.[4][5] Rutherford's isolation occurred amid a flurry of contemporaneous investigations into the composition of air during the late 18th century. Swedish chemist Carl Wilhelm Scheele independently produced a similar gas in 1772 by heating ammonium nitrate and other nitrates, though he did not fully characterize it as an element. English chemist Joseph Priestley also isolated the gas around the same time through combustion experiments, calling it "phlogisticated air," and further studied its properties by reacting it with metals to form oxides. In 1781, Henry Cavendish confirmed nitrogen's inert nature by sparking mixtures of air and oxygen over water, producing nitric acid and quantifying the proportions of atmospheric gases. Rutherford received primary credit for the discovery due to the publication of his dissertation, De aere fixo dicto aut mephitico, in 1772.[5][4]Etymology and nomenclature
The element nitrogen was first isolated in 1772 by Scottish physician Daniel Rutherford, who referred to it as "mephitic air" or "noxious air," terms reflecting its inability to support combustion or respiration, akin to earlier descriptions of "phlogisticated air" by chemists like Joseph Priestley and Henry Cavendish.[6] These early names highlighted its asphyxiating properties rather than its chemical identity.[7] In 1789, French chemist Antoine Lavoisier proposed the name azote for the gas, derived from the Greek prefix a- (meaning "without" or "not") and zōē (meaning "life"), emphasizing its role in causing suffocation by displacing oxygen.[6] This term gained traction in France and influenced names in several languages. The following year, in 1790, French chemist Jean-Antoine Chaptal coined nitrogène, which became the basis for the English "nitrogen" adopted around 1794; Chaptal's name was suggested upon recognizing the element's presence in nitric acid and nitrates.[7][6] The etymology of "nitrogen" traces to the French nitrogène, combining nitre (from Latin nitrum, referring to saltpeter or potassium nitrate) with the suffix -gène (from Greek -genēs, meaning "producing" or "begetting").[7] Thus, it literally means "nitre-forming" or "soda-producing," alluding to nitrogen's role in forming nitre compounds essential for gunpowder and fertilizers. The root nitron itself originates from ancient Greek, likely borrowed from Semitic languages denoting natural soda or salt deposits.[8] Alternative names persist in various languages, often reflecting either the azote tradition or asphyxiation. In French, Italian (azoto), Russian (azot), and Polish (azot), the term derives from Lavoisier's azote, underscoring its lifeless quality.[6] German Stickstoff, introduced by Swiss physician Christoph Girtanner in 1791, combines stick- (from ersticken, "to suffocate" or "choke") with Stoff ("substance"), directly referencing its toxic effects in pure form.[6] Similarly, Dutch stikstof follows this suffocation motif. In contrast, languages like Spanish (nitrógeno), Portuguese (nitrogênio), and Swedish (kväve, from kväva, "to choke") blend or adapt these roots.[6] In modern chemical nomenclature, nitrogen is designated by the IUPAC-approved name "nitrogen" and atomic symbol N, reflecting its position as element 7 in the periodic table.[6] The symbol N is universally used, derived straightforwardly from the English and Latin nitrogenium, ensuring consistency across scientific literature. Historical proposals like alcaligène (emphasizing alkali production) were abandoned in favor of these standardized terms.[6]Properties
Physical and atomic properties
Nitrogen is a chemical element with atomic number 7 and chemical symbol N.[1] It is a nonmetal in group 15 (pnictogens) of the periodic table, with an atomic weight ranging from 14.00643 to 14.00728 due to isotopic variation.[9] The electron configuration of the nitrogen atom is [He] 2s² 2p³, featuring three unpaired electrons in the 2p orbitals that contribute to its reactivity in bonding.[1] On the Pauling scale, nitrogen has an electronegativity of 3.04, reflecting its strong tendency to attract electrons in chemical bonds.[10] The first ionization energy is 14.5341 eV (equivalent to 1402.3 kJ/mol), indicating the energy required to remove one electron from a neutral atom in the gas phase.[11] At standard temperature and pressure, elemental nitrogen exists as a diatomic molecule (N₂), formed by a strong triple bond between two nitrogen atoms, which imparts high stability and low reactivity under ambient conditions.[12] N₂ is a colorless, odorless, and tasteless gas that is noncombustible and nontoxic, comprising about 78% of Earth's atmosphere by volume.[12] Its density at standard temperature and pressure (0°C, 1 atm) is 1.2506 g/L (or 1.2506 kg/m³).[13] Nitrogen liquefies and solidifies at very low temperatures, characteristic of its weak intermolecular forces despite the molecular triple bond. The following table summarizes key phase transition points for N₂:| Property | Value | Reference |
|---|---|---|
| Melting point | 63.3 K (-209.85°C) | Streng (1971) via NIST |
| Boiling point | 77.34 K (-195.81°C) | Jacobsen et al. (1986) via NIST |
| Triple point (T) | 63.14 K (-210.01°C) | Jacobsen et al. (1986) via NIST |
| Triple point (P) | 0.1252 bar | Jacobsen et al. (1986) via NIST |
| Critical temperature | 126.19 K (-146.96°C) | Jacobsen et al. (1986) via NIST |
| Critical pressure | 33.978 bar | Jacobsen et al. (1986) via NIST |
| Critical density | 11.18 mol/L | Jacobsen et al. (1986) via NIST |
Isotopes
Nitrogen has two stable isotopes: nitrogen-14 (¹⁴N) and nitrogen-15 (¹⁵N). Nitrogen-14 is the predominant isotope, accounting for 99.634% of naturally occurring nitrogen, while nitrogen-15 constitutes 0.366%. These abundances were precisely determined through mass spectrometry measurements of atmospheric nitrogen, establishing a standard ratio of ¹⁴N/¹⁵N at 272.0 ± 0.3.[16] Both isotopes are non-radioactive and indefinitely stable, with nitrogen-14 having a nuclear spin of 1 and nitrogen-15 a spin of 1/2, the latter enabling its use in nuclear magnetic resonance (NMR) spectroscopy for structural analysis in organic and biochemical studies.[17] Nitrogen-15 is particularly valuable as a tracer in environmental and biological research due to its stable nature and slight mass difference from nitrogen-14, which leads to isotopic fractionation in processes like nitrogen fixation and denitrification. In biogeochemical cycles, variations in the ¹⁵N/¹⁴N ratio (expressed as δ¹⁵N) help trace nutrient sources, with atmospheric nitrogen serving as the reference standard at 0‰ and terrestrial materials ranging from -20‰ to +30‰. For instance, fertilizers typically show δ¹⁵N values near 0‰, while animal manure ranges from +10‰ to +25‰, aiding in pollution source identification.[18] In addition to its stable isotopes, nitrogen has 15 known radioactive isotopes, spanning mass numbers from ¹⁰N to ²⁵N, along with one observed nuclear isomer. These isotopes are short-lived, with the longest half-life belonging to nitrogen-13 (¹³N) at 9.965 minutes, which undergoes 100% positron emission (β⁺) decay to carbon-13 (¹³C), making it suitable for positron emission tomography (PET) imaging in medical diagnostics. Other notable radioactive isotopes include nitrogen-12 (¹²N), with a half-life of 11 milliseconds and primarily β⁺ decay (98%) to carbon-12, and nitrogen-16 (¹⁶N), with a 7.13-second half-life and β⁻ decay (nearly 100%) to oxygen-16, the latter produced in nuclear reactors for coolant flow monitoring due to its high-energy gamma emission.[17] Nitrogen-17 decays primarily by β⁻ emission to oxygen-17 (branching ratio ≈99%) with a 4.173-second half-life, with a small branch (≈1%) for β⁻ decay accompanied by neutron emission to oxygen-16.[19] These isotopes are artificially produced and play roles in nuclear physics research and radiopharmaceutical applications, but their rapid decay limits natural occurrence.[20]| Isotope | Stability | Natural Abundance (%) | Half-Life | Decay Mode | Key Applications/Notes |
|---|---|---|---|---|---|
| ¹⁴N | Stable | 99.634 | - | - | Most common; basis for atomic mass of N (14.0067 u) |
| ¹⁵N | Stable | 0.366 | - | - | NMR spectroscopy; isotopic tracer in biology and ecology |
| ¹²N | Radioactive | - | 11 ms | β⁺ (98%) | Short-lived; nuclear research |
| ¹³N | Radioactive | - | 9.965 min | β⁺ (100%) | PET imaging in medicine |
| ¹⁶N | Radioactive | - | 7.13 s | β⁻ (~100%) | Reactor monitoring; gamma emitter |
| ¹⁷N | Radioactive | - | 4.173 s | β⁻ (≈99%), β⁻-n (≈1%) | Nuclear studies |
Allotropes
Nitrogen exists primarily as the diatomic molecule N₂ under standard conditions, which constitutes about 78% of Earth's atmosphere and is the most stable allotrope, characterized by a strong triple bond with a bond dissociation energy of 945 kJ/mol.[21] Atomic nitrogen, also known as monatomic or "active" nitrogen, is a highly reactive allotrope produced by electrical discharges in N₂ gas, first observed in 1910 by Lord Rayleigh as a glowing yellow afterglow known as the Lewis-Rayleigh afterglow.[22] This form consists of nitrogen atoms in the ground state (^4S) or excited states, with three unpaired electrons making it a triradical that rapidly recombines to N₂, releasing energy as light and heat; it is transient and not stable at ambient conditions.[23] Beyond diatomic and atomic forms, nitrogen forms various polynitrogen molecules, often unstable and requiring low temperatures or isolation techniques for observation. The azide radical N₃• is a short-lived species observed spectroscopically, while N₄ has been identified but not fully structurally characterized.[24] In 2025, researchers synthesized neutral hexanitrogen (N₆) via gas-phase reaction of chlorine or bromine with silver azide at room temperature, trapping it in an argon matrix at 10 K or as films in liquid nitrogen at 77 K; this linear molecule features four double bonds and one central single bond, exhibiting a high dissociation barrier but poor thermal stability, decomposing explosively to N₂ and potentially serving as a high-energy-density material.[25] Polymeric nitrogen allotropes, consisting of extended networks of single-bonded nitrogen atoms, are synthesized under extreme high-pressure conditions and represent promising energy storage materials due to their potential to release large amounts of energy upon decomposition to N₂. Cubic gauche nitrogen (cg-N), first synthesized in 2004 at pressures above 110 GPa and temperatures around 2000 K, features a diamond-like three-dimensional structure with sp³ hybridization and is metastable upon decompression to ambient pressure.[26] Other forms include black phosphorus-like nitrogen (bp-N), achieved by laser heating N₂ at 140 GPa, which adopts a layered puckered structure similar to black phosphorus and exhibits semiconducting properties.[27] Layered polymeric nitrogen (LP-N) and post-layered variants like PLP-N form at pressures between 100–240 GPa, displaying two-dimensional sheets or hexagonal layers with varying stability up to room temperature under pressure.[28] These polymeric phases are typically accessed using diamond anvil cells and confirmed via X-ray diffraction, highlighting nitrogen's ability to mimic carbon's allotropy under compression.[29]Chemistry
Reactivity and bonding
Nitrogen exists primarily as the diatomic molecule N₂, characterized by a strong triple bond (N≡N) with a bond dissociation energy of 941 kJ/mol, which renders it highly stable and contributes to its low chemical reactivity under ambient conditions.[30] This triple bond arises from the overlap of one σ orbital (from sp hybridization) and two π orbitals (from unhybridized p orbitals), allowing each nitrogen atom to achieve an octet configuration while sharing six electrons.[31] The high bond energy—nearly twice that of the O=O double bond in O₂ (498 kJ/mol)—stems from nitrogen's small atomic size and effective orbital overlap, making bond cleavage endothermic and kinetically unfavorable without activation.[31] Due to this stability, N₂ exhibits minimal reactivity at room temperature, interacting only with highly electropositive metals such as lithium to form nitrides like Li₃N, or with certain biological enzymes in nitrogen-fixing bacteria that employ metal cofactors to weaken the bond.[32] Enhanced reactivity occurs under extreme conditions, including high temperatures (e.g., above 1000°C for combination with oxygen to yield NO), high pressures, or catalysis; the industrial Haber-Bosch process, for instance, uses iron-based catalysts at 200–300 atm and 400–500°C to facilitate N₂ + 3H₂ → 2NH₃.[31] In contrast, atomic nitrogen or activated N₂ species, such as those generated in electrical discharges, display high reactivity, forming azides or inserting into C–H bonds.[32] In bonding, nitrogen's five valence electrons enable versatile hybridization (sp, sp², sp³) and multiple bond formation, favoring π-bonding with second-period elements like carbon and oxygen due to similar orbital sizes and energies.[31] Common motifs include triple bonds in cyanides (e.g., H–C≡N), double bonds in imines (R₂C=NR), and single bonds in amines (R–NH₂), with catenation allowing chains like hydrazine (H₂N–NH₂).[31] Nitrogen's electronegativity (3.04 on the Pauling scale) promotes polar covalent bonds, where it often acts as a Lewis base via its lone pair, as in ammonia coordinating to metal centers.[31] Nitrogen spans oxidation states from −3 (e.g., in NH₃ and N³⁻ nitrides) to +5 (e.g., in NO₃⁻ and HNO₃), reflecting its ability to gain three electrons for reduction or lose five for oxidation, which underpins its redox versatility in compounds.[31] Intermediate states include +2 in NO (with a formal double bond and unpaired electron, making it paramagnetic) and +4 in NO₂ (dimerizing to N₂O₄ via N–N single bond).[31] This range facilitates reactions like the oxidation of NH₃ to HNO₃ in the Ostwald process, where nitrogen cycles through multiple states.[31]Inorganic compounds
Inorganic nitrogen compounds encompass a diverse array of species reflecting nitrogen's variable oxidation states from −3 to +5, enabling it to form stable bonds with metals, nonmetals, hydrogen, and oxygen. These compounds are pivotal in industrial processes, agriculture, and environmental chemistry, with ammonia and nitric acid ranking among the most produced chemicals globally.[33][34][31] Nitrides represent compounds where nitrogen adopts the −3 oxidation state, typically forming with alkali and alkaline earth metals. Lithium nitride (Li₃N) is synthesized by direct reaction of lithium metal with dinitrogen at ambient temperature, yielding a solid that reacts vigorously with water to produce ammonia and lithium hydroxide.[2][31] Transition metal nitrides, such as titanium nitride (TiN), exhibit high hardness and melting points, often prepared by heating metals in ammonia or nitrogen atmospheres, and find applications in coatings and ceramics due to their refractory properties.[2] Hydrides of nitrogen, primarily ammonia (NH₃) and hydrazine (N₂H₄), are key binary compounds with hydrogen. Ammonia, the most abundant inorganic nitrogen compound industrially, is produced via the Haber-Bosch process, which combines dinitrogen and dihydrogen at 200–300 atm and 400–500°C over an iron catalyst, achieving equilibrium yields optimized by Le Châtelier's principle.[2][31] It acts as a weak base in aqueous solution, forming the ammonium ion (NH₄⁺) with pK_a ≈ 9.25, and serves as a precursor for fertilizers and explosives. Hydrazine, prepared by oxidation of ammonia with sodium hypochlorite, is a high-energy liquid used in rocket propulsion, combusting exothermically with oxygen (ΔH = −622 kJ/mol) to yield dinitrogen and water.[2][31][34][35] Nitrogen oxides constitute a series of gaseous compounds with oxidation states ranging from +1 to +5, many of which play roles in atmospheric and biological processes. Nitrous oxide (N₂O), or laughing gas, is generated by thermal decomposition of ammonium nitrate at 170–200°C and serves as an anesthetic due to its inertness under physiological conditions, though it contributes to stratospheric ozone depletion.[2][31] Nitric oxide (NO), an odd-electron radical formed in combustion or by reduction of nitric acid with copper, acts as a signaling molecule in mammals for vasodilation and neurotransmission, while reacting with dioxygen to form nitrogen dioxide (NO₂).[2][31][34] NO₂, a toxic brown gas that dimerizes to N₂O₄ at lower temperatures, is a key pollutant from engine exhausts, oxidizing to nitrogen pentoxide (N₂O₅) in air, which hydrolyzes to nitric acid.[31] Oxoacids and their anions, such as nitric acid (HNO₃) and nitrous acid (HNO₂), exemplify nitrogen in higher oxidation states. Nitric acid, produced industrially by the Ostwald process—oxidizing ammonia to NO over platinum, then to NO₂, followed by absorption in water—is a strong, fuming acid used in nitrate fertilizer production and explosives like ammonium nitrate.[33][31] It exhibits +5 oxidation state and acts as a powerful oxidizer, decomposing metals to release NO or NO₂ depending on concentration. Nitrite ions (NO₂⁻) from nitrous acid serve as preservatives in food, while nitrate ions (NO₃⁻) are essential in fertilizers but can cause eutrophication in waterways.[2][34][36] Nitrogen halides, though less stable than fluorides, include nitrogen trifluoride (NF₃), synthesized from ammonia and fluorine over copper catalyst, which is thermodynamically stable and used in plasma etching due to its +3 oxidation state.[34][31] Nitrogen trichloride (NCl₃), formed from ammonia and chlorine, is highly explosive and decomposes to dinitrogen and chlorine. Azides, containing the N₃⁻ ion, are energetic materials; sodium azide (NaN₃) is prepared by reacting sodium amide with nitrous oxide and detonates to release dinitrogen, powering automobile airbags.[2][34] These compounds highlight nitrogen's versatility, though many require careful handling due to reactivity.[34]Organic nitrogen compounds
Organic nitrogen compounds are a diverse class of substances in which nitrogen atoms are covalently bonded to carbon atoms, forming part of the molecular backbone or functional groups. These compounds play crucial roles in organic synthesis, pharmaceuticals, dyes, and biological systems, with nitrogen's ability to form up to four bonds enabling a wide range of structures and reactivities.[37] They are classified based on the functional group involving nitrogen, including amines, amides, nitriles, nitro compounds, imines, and nitrogen-containing heterocycles, among others.[38] Amines are derivatives of ammonia (NH₃) where one or more hydrogen atoms are replaced by organic substituents such as alkyl or aryl groups. They are categorized as primary (R-NH₂), secondary (R₂NH), or tertiary (R₃N), with quaternary ammonium salts (R₄N⁺X⁻) formed by further alkylation. Amines exhibit basicity due to the lone pair on nitrogen, which accepts a proton, and their boiling points increase with hydrogen bonding capability—primary and secondary amines form hydrogen bonds, while tertiary ones do not. Common examples include methylamine (CH₃NH₂), used in chemical synthesis, and aniline (C₆H₅NH₂), a precursor for dyes and pharmaceuticals; they often have a fishy odor and are synthesized industrially from alcohols and ammonia.[38] Amides feature a nitrogen atom bonded to a carbonyl group (R-C(O)-NR₂), resulting from the condensation of carboxylic acids and amines, releasing water. This functional group is planar due to resonance between the nitrogen lone pair and the carbonyl π-bond, reducing basicity compared to amines and imparting stability. Primary amides (R-C(O)NH₂) like acetamide are soluble in water via hydrogen bonding, while secondary and tertiary amides appear in peptides and synthetic polymers. Amides are key in biochemistry as peptide bonds in proteins and in materials like nylon, formed from diamines and dicarboxylic acids.[38][39] Nitriles, or cyanides (R-C≡N), contain a carbon-nitrogen triple bond, making them versatile intermediates in organic synthesis for converting to carboxylic acids, esters, or amines via hydrolysis or reduction. The cyano group is electron-withdrawing, influencing reactivity in adjacent positions, and nitriles are typically liquids or low-melting solids with high dipole moments. Acetonitrile (CH₃CN) serves as a common solvent, while adiponitrile is used in nylon production; they are prepared from halides and cyanide ions or dehydration of amides.[40] Nitro compounds (R-NO₂) consist of a nitro group attached to carbon, often via electrophilic aromatic substitution on arenes or direct nitration of alkanes. The nitro group is strongly electron-withdrawing, activating ortho/para positions for further electrophilic attack in aromatics and stabilizing carbanions in aliphatic cases. Nitrobenzene (C₆H₅NO₂) is a major industrial compound, reduced to aniline for dyes and drugs, while nitromethane (CH₃NO₂) acts as a solvent and explosive precursor; these compounds are typically pale yellow liquids or solids with high boiling points.[41] Imines (R₂C=NR') arise from condensation of carbonyl compounds with primary amines, featuring a C=N double bond that imparts rigidity and is key in Schiff bases for coordination chemistry. They are less stable than carbonyls and hydrolyze under acidic or basic conditions, with examples like benzaldimine used in synthetic routes to heterocycles.[39] Nitrogen-containing heterocyclic compounds integrate nitrogen into ring structures, such as five- or six-membered rings, exhibiting aromaticity and unique reactivities. Pyrrole (five-membered, NH in ring) contributes to porphyrins in heme, while pyridine (six-membered, N replacing CH) behaves like benzene but with basic nitrogen for nucleophilic substitutions. These are foundational in alkaloids, vitamins (e.g., nicotine), and pharmaceuticals, with synthesis often involving cyclization reactions.[38][37] Other notable classes include azo compounds (R-N=N-R), valued for dyes due to visible light absorption, and hydrazines (R-NH-NH₂), used in rocket fuels and as reducing agents. Overall, the reactivity of organic nitrogen compounds stems from nitrogen's electronegativity and lone pair, enabling applications across chemistry and biology.[39][42]Occurrence
In the universe and Earth's crust
Nitrogen is among the most abundant elements in the universe, ranking sixth overall by atomic number fraction, with a cosmic abundance of approximately $7.5 \times 10^{-5} atoms relative to hydrogen (or \log \epsilon(\mathrm{N}) = 7.83 \pm 0.05). This places it at roughly 0.007% of the total number of atoms in the cosmos, behind hydrogen, helium, oxygen, carbon, and neon. The element is primarily synthesized through the carbon-nitrogen-oxygen (CNO) cycle, a proton-capture process occurring in the cores of main-sequence stars more massive than about 1.3 solar masses, where carbon acts as a catalyst to fuse hydrogen into helium, producing nitrogen as a byproduct.[43][44] In the interstellar medium, nitrogen exists predominantly as molecular N₂, with detections confirmed via ultraviolet spectroscopy, though atomic and ionic forms are also present in diffuse regions. Its distribution reflects stellar nucleosynthesis over cosmic time, with higher abundances observed in metal-rich environments due to contributions from asymptotic giant branch (AGB) stars and supernovae. Observations of distant galaxies indicate that nitrogen enrichment correlates with star formation history, serving as a tracer for galactic chemical evolution.[45] In contrast to its cosmic prevalence, nitrogen is scarce in Earth's crust, comprising only about 83 ppm by weight in the upper continental crust and around 50–74 ppm in the bulk crust. This low concentration arises because nitrogen is volatile and largely partitioned into the atmosphere and hydrosphere during planetary differentiation, with minimal incorporation into silicate minerals. It occurs mainly as nitrate minerals (e.g., in evaporites), ammonium ions adsorbed onto clays or substituted in micas and feldspars, and trace nitrides in meteoritic materials, though organic-bound forms dominate in sedimentary rocks. Subduction and metamorphic processes can release crustal nitrogen back to the mantle or atmosphere, influencing the global nitrogen cycle.[46][47]In the atmosphere
Nitrogen constitutes approximately 78% of Earth's atmosphere by volume in dry air, making it the most abundant gas and a primary component influencing overall atmospheric pressure and density. This dominance arises from the stability of its diatomic form, N₂, which accounts for nearly all atmospheric nitrogen due to the molecule's strong triple bond, rendering it chemically inert under typical conditions.[48][49][50] The presence of N₂ serves a critical role in maintaining atmospheric balance by diluting oxygen concentrations to about 21%, which prevents uncontrolled combustion and rapid burning at the Earth's surface. Without this buffering effect, the higher oxygen levels would accelerate fire propagation and oxidation processes, potentially altering ecosystems and climate dynamics. N₂'s inertness also contributes to the long-term stability of the atmosphere, as it resists reaction with other gases except under high-energy conditions like lightning or in the upper atmosphere.[51][48] While N₂ overwhelmingly predominates, trace reactive nitrogen compounds exist in the atmosphere at concentrations below 0.01%, including nitric oxide (NO), nitrogen dioxide (NO₂), nitrous oxide (N₂O), and ammonia (NH₃). These species originate from natural sources such as soil emissions, wildfires, and microbial activity, as well as human activities like fossil fuel combustion and agriculture. They participate in key atmospheric processes, including the formation of tropospheric ozone, aerosol production, and nutrient deposition to ecosystems, though their low abundances limit direct impacts on bulk atmospheric composition.[52][53]In living organisms
Nitrogen is an essential constituent of all known living organisms, where it ranks as the fourth most abundant element in cellular biomass after carbon, hydrogen, and oxygen.[54] It occurs primarily in organic forms integrated into vital biomolecules, and its abundance varies by organism type and environmental conditions. In the terrestrial biosphere, the total nitrogen reservoir associated with living systems is estimated at around 135 Gt, though much of this is cycled through non-living organic matter; live plant biomass alone holds approximately 5% of this pool, or about 6.75 Gt of nitrogen. In plants, nitrogen constitutes 1–5% of the total dry matter, making it the most abundant mineral nutrient after carbon, hydrogen, and oxygen.[55] This proportion is highest in metabolically active tissues like leaves, where it supports photosynthesis and growth, and lower in woody parts. Healthy above-ground plant tissues often contain 3–4% nitrogen by dry weight, reflecting its concentration in chlorophyll and enzymes.[56] Leguminous plants can achieve higher levels through symbiotic nitrogen fixation, enhancing overall biomass nitrogen content. In animals, nitrogen typically accounts for about 3% of total body mass, as seen in humans, where it is concentrated in muscle proteins and other tissues.[57] On a dry weight basis, this rises to roughly 7–12%, depending on the species and tissue type, due to the exclusion of water. Microbial biomass, including bacteria and fungi, exhibits even higher nitrogen concentrations, often 10–14% of dry weight, driven by their protein-rich composition. For example, in the bacterium Escherichia coli, nitrogen comprises approximately 12% of dry biomass under standard growth conditions.[58] Globally, vegetation dominates the nitrogen inventory in living organisms, while microbial communities contribute substantially in soils (up to several Gt N in aggregate) and marine environments.Biological role
Nitrogen fixation
Biological nitrogen fixation (BNF) is the microbial process that converts atmospheric dinitrogen (N₂) into bioavailable forms, primarily ammonia (NH₃), which plants and other organisms can assimilate for growth.[59] This process is essential for maintaining global nitrogen cycles, as it provides the primary natural input of fixed nitrogen, estimated at approximately 128 teragrams (Tg) of nitrogen per year in natural terrestrial ecosystems and 60 Tg in agricultural systems.[60] Without BNF, ecosystems would lack sufficient nitrogen for protein synthesis and other vital functions, leading to severe limitations in primary productivity.[61] The core mechanism of BNF is catalyzed by the nitrogenase enzyme complex, a metalloenzyme consisting of the Fe protein (also called nitrogenase reductase) and the MoFe protein, encoded by nif genes.[62] Nitrogenase reduces N₂ to NH₃ using electrons from ferredoxin or flavodoxin and requires 16 ATP molecules per N₂ molecule fixed, making it highly energy-intensive.[59] The enzyme is extremely sensitive to oxygen, which inactivates it; thus, diazotrophs (nitrogen-fixing microorganisms) employ protective strategies such as spatial separation in heterocysts (in cyanobacteria like Anabaena), rapid respiration in free-living bacteria like Azotobacter, or isolation within plant root nodules in symbiotic associations.[59] Alternative nitrogenases containing vanadium (V-nitrogenase) or iron only (Fe-nitrogenase) exist but are less efficient and active under specific metal-limited conditions.[59] Genetic regulation of BNF ensures expression only under nitrogen-limiting conditions to conserve energy. In free-living diazotrophs like Azotobacter vinelandii, the enhancer-binding protein NifA activates transcription of nif genes via σ⁵⁴-dependent promoters, while the inhibitory protein NifL, a flavoprotein, senses redox status and ammonium levels to repress NifA activity.[62] In symbiotic systems, such as Rhizobium with legumes, oxygen sensing occurs via the FixL/FixJ two-component system, where low oxygen activates NifA to initiate nodule-specific nif gene expression.[62] These regulatory cascades integrate signals from nitrogen availability, oxygen, and plant host factors to coordinate fixation.[62] BNF occurs through three main types: free-living, associative, and symbiotic. Free-living diazotrophs, including aerobic Azotobacter and anaerobic Clostridium, fix nitrogen in soil, contributing 20–30 kg N ha⁻¹ yr⁻¹, though rates vary widely (3–51 kg N ha⁻¹ per crop).[59] Associative symbiosis involves endophytic bacteria like Azospirillum in cereal roots, providing modest inputs (up to 82% of plant nitrogen derived from air in some maize studies).[60] Symbiotic fixation is most efficient, particularly in legumes where Rhizobium or Bradyrhizobium form nodules on roots, fixing 20–300 kg N ha⁻¹ yr⁻¹; for example, soybeans globally fix about 24.8 Tg N annually, representing 70% of legume contributions.[59] Other symbioses include Frankia with actinorhizal plants and cyanobacteria like Nostoc in lichens or Azolla in rice paddies (20–146 kg N ha⁻¹).[60] In agriculture, BNF supports sustainable practices by reducing reliance on synthetic fertilizers; grain legumes alone provide 35.5 Tg N globally per year, enhancing soil fertility for subsequent crops through residual nitrogen (30–50 kg N ha⁻¹ leaked to soil).[63] Ecologically, BNF drives primary production in diverse habitats, including inland and coastal waters where it fixes 40 Tg N yr⁻¹, accounting for 15% of total global marine and terrestrial fixation.[61] Efforts to engineer BNF into non-leguminous cereals, such as through microbial inoculation or genetic transfer of nif genes, aim to boost yields while minimizing environmental impacts like eutrophication from fertilizer runoff.[60]Role in biomolecules
Nitrogen is an essential element in numerous biomolecules, forming the structural backbone of key macromolecules that underpin cellular function, growth, and heredity in all living organisms. It constitutes a critical component of amino acids, nucleic acids, and other vital compounds, enabling the synthesis of proteins, genetic material, and energy carriers. Without biologically available nitrogen, these molecules could not form, limiting the development and maintenance of life.[64][49] In proteins, nitrogen is integral to the amino group (-NH₂) present in every amino acid, the fundamental building blocks of these macromolecules. Proteins, which can comprise up to 55% of a prokaryote's dry weight, serve diverse roles as enzymes catalyzing metabolic reactions, structural components like collagen, and signaling molecules such as hormones. The incorporation of nitrogen into amino acids occurs through assimilation processes where ammonia (NH₃) or nitrate is converted into glutamine or glutamate, precursors for all other amino acids. This nitrogen-dependent synthesis is vital for protein folding and function, directly influencing organismal physiology and adaptation.[65][66] Nucleic acids, including DNA and RNA, rely on nitrogenous bases—purines (adenine, guanine) and pyrimidines (cytosine, thymine, uracil)—which contain multiple nitrogen atoms within their heterocyclic rings. These bases, accounting for about 23% of prokaryotic dry weight in nucleotide forms, enable the storage, replication, and expression of genetic information. Nitrogen's presence in these structures facilitates base pairing and hydrogen bonding, essential for the double-helix stability of DNA and the functional versatility of RNA in processes like transcription and translation. The synthesis of these bases begins with nitrogen-rich precursors like aspartate and glutamine, highlighting nitrogen's irreplaceable role in heredity.[64][66][65] Beyond proteins and nucleic acids, nitrogen features in other biomolecules such as chlorophyll, where it forms the porphyrin ring central to photosynthesis, allowing plants and algae to convert light energy into chemical energy. It is also a key element in ATP (adenosine triphosphate), the primary energy currency of cells, and in amino sugars like chitin, which provide structural support in fungal cell walls and arthropod exoskeletons. These roles underscore nitrogen's versatility in supporting energy metabolism, photosynthesis, and extracellular matrices across ecosystems.[64][65]The nitrogen cycle
The nitrogen cycle is a biogeochemical process through which nitrogen is converted among its multiple chemical forms—including dinitrogen gas (N₂), ammonia (NH₃), ammonium (NH₄⁺), nitrite (NO₂⁻), nitrate (NO₃⁻), and organic nitrogen compounds—as it circulates among the atmosphere, terrestrial and aquatic ecosystems, and living organisms.[64] This cycle is fundamental to sustaining life on Earth, as nitrogen is an essential element in amino acids, proteins, nucleic acids, and other biomolecules required for growth and reproduction.[67] Microorganisms, particularly bacteria and archaea, drive most transformations, while abiotic processes like lightning contribute smaller amounts.[64] The cycle begins with nitrogen fixation, the conversion of atmospheric N₂ into biologically available ammonia (NH₃) or ammonium (NH₄⁺). This energy-intensive process requires 16 ATP molecules and eight electrons per N₂ molecule and is primarily biological, performed by diazotrophic prokaryotes such as symbiotic bacteria (Rhizobium spp. in legume root nodules) and free-living bacteria (Azotobacter spp.).[64] Abiotic fixation occurs via lightning or industrial processes like the Haber-Bosch method, but biological fixation accounts for the majority of natural inputs, estimated at around 100–200 teragrams of nitrogen per year globally.[67] Once fixed, ammonia is rapidly assimilated by plants and microbes into organic forms, such as amino acids, through nitrogen assimilation.[64] Organic nitrogen from dead organisms, waste, or plant residues is then mineralized back into ammonium via ammonification (or mineralization), a decomposition process carried out by diverse heterotrophic bacteria and fungi that break down proteins and other nitrogenous compounds.[64] In aerobic soils and waters, ammonium is oxidized to nitrate in a two-step nitrification process: first to nitrite by ammonia-oxidizing bacteria (Nitrosomonas spp.) or archaea (Nitrosopumilus maritimus), and then to nitrate by nitrite-oxidizing bacteria (Nitrospira spp.).[64] This autotrophic process supports plant uptake of nitrate as a primary nitrogen source.[67] Under anaerobic conditions, such as waterlogged soils or sediments, nitrate is reduced back to N₂ through denitrification, a respiratory process performed by facultative anaerobes like Pseudomonas and Paracoccus spp., which use nitrate as an electron acceptor instead of oxygen.[64] This step releases dinitrogen gas (N₂) and nitrous oxide (N₂O), a potent greenhouse gas, closing the cycle by returning nitrogen to the atmosphere.[67] An additional anaerobic pathway, anaerobic ammonium oxidation (anammox), directly converts ammonium and nitrite to N₂ using specialized bacteria from the phylum Planctomycetes (Candidatus Brocadia anammoxidans), contributing significantly to nitrogen loss in marine and freshwater systems—up to 50% of oceanic N₂ production in some oxygen minimum zones.[64] These interconnected processes maintain a dynamic balance in the reactive nitrogen pool, with global fluxes dominated by microbial activity; for instance, biological fixation vastly exceeds abiotic sources by a factor of 10 or more.[67] Human activities, including fertilizer production and fossil fuel combustion, have doubled reactive nitrogen inputs since the mid-20th century, accelerating the cycle and leading to imbalances such as excess nitrate in ecosystems.[64]Production
Industrial processes
The primary industrial production of nitrogen gas (N₂) occurs through the separation of air, which consists of approximately 78% nitrogen by volume. This is achieved via three main processes: cryogenic distillation, pressure swing adsorption (PSA), and membrane permeation, each suited to different scales and purity requirements. Cryogenic methods dominate large-scale production, while PSA and membrane techniques enable on-site generation for medium to small volumes. These processes supply nitrogen for applications in chemicals, food preservation, electronics, and metallurgy, primarily from air separation units (ASUs).[68] Cryogenic air separation, the most established method, involves compressing atmospheric air to 5–10 bar, cooling it to remove impurities like water vapor and carbon dioxide via refrigeration and molecular sieves, and then liquefying it at around -196°C using the Joule-Thomson effect and expansion turbines. The liquefied air enters a double distillation column where nitrogen, with its lower boiling point (-196°C), vaporizes and rises to the top, while oxygen (-183°C) and other components remain liquid or distill separately; argon can be extracted as a byproduct in advanced setups. This process, pioneered in the late 19th century and commercialized in the early 20th, achieves ultra-high purity levels of 99.999% or greater and is energy-efficient for large plants producing thousands of tons per day, though it requires significant capital investment and is less viable for small-scale operations.[69][70] Pressure swing adsorption (PSA) provides a non-cryogenic alternative by passing compressed air (typically 7–10 bar) through beds of carbon molecular sieves or zeolites that preferentially adsorb oxygen and water vapor, allowing nitrogen to pass through. The process cycles between high-pressure adsorption (where oxygen is captured) and low-pressure desorption (purging the adsorbent with a portion of the product nitrogen), often using two or more beds in parallel for continuous operation. Developed in the mid-20th century and refined since the 1980s, PSA systems yield nitrogen purities of 95–99.999%, with recoveries around 70–90%, and are cost-effective for medium-scale on-site production (up to 100 tons per day) due to lower energy use (about 0.3–0.5 kWh/Nm³) compared to cryogenic methods, though they consume more air per unit of nitrogen.[71][72] Membrane separation, another non-cryogenic technique, employs bundles of hollow polymeric fibers (e.g., polyimide or polysulfone) that selectively permeate faster-diffusing gases like oxygen and carbon dioxide through their walls under pressure (8–15 bar), retaining slower-permeating nitrogen on the high-pressure side. The process, commercialized in the 1980s, requires pre-treatment of air to remove oil and particulates, and a single-stage setup typically produces nitrogen at 95–99.5% purity with 40–50% recovery, while multi-stage configurations can exceed 99%. It is compact, modular, and ideal for small-scale, portable applications (e.g., 1–50 Nm³/h) with low maintenance, but purity is generally lower than cryogenic or PSA methods, limiting its use in high-precision industries. Advances in membrane materials continue to improve selectivity and durability.[68] In addition to direct N₂ production, the Haber-Bosch process indirectly utilizes nitrogen by reacting it with hydrogen to form ammonia (NH₃) on an iron-based catalyst at 200–300 bar and 400–500°C, accounting for over 90% of global fixed-nitrogen output (approximately 185 million tons of ammonia annually as of 2023).[73] Nitrogen for this process is sourced from air separation, primarily cryogenic ASUs integrated with ammonia plants to supply high-purity N₂. This method revolutionized agriculture but is energy-intensive, consuming about 1–2% of global energy.[68]Laboratory preparation
In the laboratory, nitrogen gas (N₂) is commonly prepared by the thermal decomposition of ammonium nitrite (NH₄NO₂), which is generated in situ from an equimolar mixture of ammonium chloride (NH₄Cl) and sodium nitrite (NaNO₂) to avoid handling the unstable pure compound.[74] The reaction proceeds as follows: \mathrm{NH_4Cl}(s) + \mathrm{NaNO_2}(s) \rightarrow \mathrm{NaCl}(s) + \mathrm{N_2}(g) + 2\mathrm{H_2O}(g) This method yields nitrogen gas quantitatively at temperatures around 240–260°C, with the evolved gas collected by downward displacement of water due to its low solubility (approximately 0.02 g/100 mL at 20°C).[74][75] The procedure involves dissolving equimolar amounts of NH₄Cl and NaNO₂ in a minimal volume of water in a round-bottom flask, then gently heating the mixture while avoiding excessive temperatures to prevent side reactions that could produce nitric oxide (NO) or ammonia (NH₃).[75] The resulting nitrogen gas may contain trace impurities such as water vapor, NH₃, or oxygen (O₂), which are removed by passing the gas through concentrated sulfuric acid to absorb moisture and basic gases, followed by a tube of heated copper turnings (at about 400°C) to reduce any O₂ to CuO.[76] The purified gas is then dried over phosphorus pentoxide (P₄O₁₀) or calcium chloride (CaCl₂) for complete dehydration.[77] Alternative laboratory methods include the oxidation of ammonia with sodium hypobromite, following the equation: $3\mathrm{NaOBr} + 2\mathrm{NH_3} \rightarrow \mathrm{N_2} + 3\mathrm{NaBr} + 3\mathrm{H_2O} This reaction is carried out at room temperature by adding bromine water to concentrated ammonia solution, producing nitrogen gas that is similarly purified.[78] Another approach involves the thermal decomposition of metal azides, such as barium azide (Ba(N₃)₂), at 150–200°C: \mathrm{Ba(N_3)_2}(s) \rightarrow \mathrm{Ba}(s) + 3\mathrm{N_2}(g) However, this method requires careful control to manage the metallic residue and is less common for general laboratory use due to safety concerns with azides.[74] These techniques ensure high-purity nitrogen suitable for experiments, contrasting with industrial fractional distillation of liquefied air, which is not feasible on a small scale.[74]Applications
Agricultural and fertilizer uses
Nitrogen is an essential macronutrient for plant growth, forming the building blocks of amino acids, proteins, nucleic acids, and chlorophyll, which are critical for photosynthesis and overall crop productivity.[79] In agriculture, nitrogen fertilizers address soil deficiencies caused by intensive cropping, enabling higher yields and supporting global food security.[80] Without supplemental nitrogen, natural soil supplies would limit crop production to levels insufficient for the world's population.[81] The widespread use of nitrogen fertilizers began with the Haber-Bosch process, invented in 1909 by Fritz Haber and Carl Bosch, which synthesizes ammonia from atmospheric nitrogen and hydrogen (primarily from natural gas) under high pressure and temperature using an iron catalyst.[81] This industrial method scaled up during World War I for explosives and post-war for fertilizers, increasing global nitrogen application from 1.3 million metric tons in 1930 to about 80 million metric tons by 1988, a roughly 60-fold rise. As of 2022, global nitrogen fertilizer consumption had reached approximately 109 million metric tons of nutrient.[80][82] In the United States, synthetic nitrogen fertilizers transformed corn production, with usage stabilizing around 10-12 million metric tons annually since the 1970s, primarily as anhydrous ammonia, urea, and urea-ammonium nitrate (UAN) solutions.[81] Nitrogen fertilizers are categorized by their chemical form, each suited to specific soil types, crops, and application timings to minimize losses from leaching, volatilization, or denitrification.[83] The table below summarizes common types, their nitrogen content, and agricultural applications:| Fertilizer Type | Nitrogen Content (%) | Key Uses and Characteristics |
|---|---|---|
| Anhydrous Ammonia | 82 | Injected into soil for fall plowdown in corn on medium- to heavy-textured soils; high efficiency but requires specialized equipment to avoid volatilization.[83] |
| Urea | 46 | Surface-applied or incorporated for top-dressing wheat and pastures in late winter; converts to ammonium via hydrolysis, but prone to ammonia loss if not incorporated above 50°F (10°C).[83] |
| Ammonium Nitrate | 34 | Top-dressing for wheat or spring plowdown for corn; provides both ammonium and nitrate forms for quick availability, though it poses explosion risks if mishandled.[83] |
| Urea-Ammonium Nitrate (UAN) Solutions | 28-32 | Sidedressing for corn via irrigation or spray; versatile liquid form for banded application, reducing overall usage rates.[83] |
| Ammonium Sulfate | 21 | Fall plowdown or top-dressing for sulfur-deficient soils in pastures and wheat; acidifies soil slightly and resists volatilization on neutral to alkaline soils.[83] |
| Calcium Nitrate | 15.5 | Supplemental for high-pH or calcium-deficient soils in vegetables and fruits; nitrate form minimizes acidification but increases leaching risk on sandy soils.[83] |