Fact-checked by Grok 2 weeks ago

Nitrogen cycle

The nitrogen cycle is a biogeochemical process through which nitrogen circulates among the atmosphere, terrestrial and aquatic ecosystems, and living organisms by undergoing transformations between inert dinitrogen gas (N₂) and reactive forms such as ammonia (NH₃), nitrates (NO₃⁻), and organic compounds. Primarily mediated by microorganisms, the cycle enables the fixation of atmospheric nitrogen—comprising about 78% of Earth's atmosphere—into bioavailable nutrients essential for synthesizing amino acids, proteins, nucleic acids, and chlorophyll, thereby supporting primary production and all higher trophic levels. Key transformations include , where prokaryotes like diazotrophs convert N₂ to NH₃; , in which plants incorporate or nitrates into ; mineralization, releasing from via decomposers; , oxidizing to nitrates by specialized ; and , reducing nitrates back to N₂ under conditions. These microbial-driven steps maintain nitrogen's availability despite its abundance in inaccessible forms, with abiotic contributions from and wildfires playing minor roles. Anthropogenic perturbations, notably from Haber-Bosch process fertilizer production and fossil fuel combustion, have roughly doubled global reactive nitrogen creation since the Industrial Revolution, amplifying fluxes and contributing to environmental issues such as soil acidification, water eutrophication, and greenhouse gas emissions via nitrous oxide (N₂O). While enhancing agricultural yields and human food security, these alterations underscore the cycle's sensitivity to external inputs, with ongoing research emphasizing microbial community dynamics and potential feedbacks in a changing climate.

Historical Development

Early Scientific Discoveries

Daniel Rutherford first isolated nitrogen gas in 1772 while conducting experiments at the University of Edinburgh, removing oxygen and carbon dioxide from air via combustion and absorption with limewater, leaving a residue that neither supported combustion nor respiration. This residual gas, termed "noxious air" or "mephitic air" by Rutherford, comprised the fixed portion of air unresponsive to life-sustaining processes, as detailed in his doctoral dissertation. Independent isolations followed shortly thereafter: Swedish chemist Carl Wilhelm Scheele obtained the gas around 1772 by heating nitrogenous compounds and air mixtures, recognizing it as a distinct atmospheric component alongside "fire air" (oxygen). Henry Cavendish similarly produced nitrogen in the mid-1770s by passing air over heated copper to deoxygenate it, then removing resulting copper oxide, confirming the gas's prevalence in air through quantitative pneumatic trough measurements. French chemist Antoine Lavoisier, building on these findings in the late 1770s, named the element azote—from Greek roots meaning "without life"—due to its inability to sustain animal respiration or flame, while establishing through eudiometry that it constituted approximately four-fifths (about 78-80%) of dry atmospheric air. Cavendish's subsequent experiments in the 1780s further elucidated nitrogen's chemical behavior, demonstrating that electric sparks passed through mixtures of the gas and oxygen over water produced nitric acid, revealing its capacity to form reactive oxides despite atmospheric inertness. These observations, grounded in precise volumetric analyses, distinguished nitrogen's diatomic N₂ form as non-reactive under ordinary conditions—failing to combust or dissolve readily—yet prone to combination under energetic conditions like electrical discharge or high heat, laying empirical groundwork for later reactivity studies. Early respiration trials, where small animals suffocated in pure nitrogen atmospheres within minutes, corroborated its asphyxiant properties, contrasting sharply with oxygen's vital role.

Recognition of Biological and Agricultural Roles

In 1840, Justus von Liebig articulated the law of the minimum, asserting that crop growth is constrained by the most deficient essential nutrient rather than total resource availability, with field trials revealing nitrogen as a primary limiting factor for plant yields due to its scarcity in many soils. Liebig's empirical observations, drawn from yield responses in nutrient addition experiments, highlighted nitrogen's critical role in organic compound formation within plants, shifting focus from earlier views of atmospheric origins to soil-based deficiencies. Commencing in 1843, John Bennet Lawes and Joseph Henry conducted systematic long-term trials at Rothamsted Experimental , isolating 's effects on and other crops through controlled manuring plots that excluded other variables. Their from the 1840s and 1850s showed that unfertilized plots yielded markedly lower harvests, attributing gains to via in , while legume-inclusive rotations—such as under —boosted subsequent yields by 20-50% through retention via associations, without invoking atmospheric fixation . These experiments quantified 's depletion under continuous cropping, with losses exceeding replenishment in intensive systems. Pre-industrial agriculture relied on limited natural nitrogen inputs, estimated at 40-100 Tg N per year from terrestrial biological fixation, far below demands of expanded cultivation and contributing to chronic soil exhaustion that capped yields at 1-2 tons per hectare for cereals, amplifying vulnerability to crop failures and famine during climatic stresses. Global soil analyses in this era confirmed patchy nitrogen distribution, necessitating practices like fallowing and manure recycling to avert widespread deficiencies, though these proved insufficient for population pressures evident in 19th-century yield stagnation.

Industrial Advancements and Conceptual Integration

The Haber-Bosch process, developed by Fritz Haber and Carl Bosch, achieved the first laboratory demonstration of atmospheric nitrogen fixation into ammonia in 1909, with industrial-scale production commencing at a BASF plant in Oppau, Germany, in 1913. This breakthrough enabled the synthesis of approximately 100 teragrams (Tg) of nitrogen annually for fertilizers by the late 20th century, a flux comparable to natural biological fixation rates of 100-200 Tg N per year and pivotal in expanding global food production beyond pre-industrial constraints. Conceptualization of the nitrogen cycle advanced in the late 19th and early 20th centuries through microbiological isolations that linked key transformations. Sergei Winogradsky isolated in 1890, demonstrating the two-step oxidation of to nitrite and then via distinct organisms, and . Martinus identified symbiotic nitrogen-fixing in legume root nodules around 1888-1901, establishing biological fixation as a microbial converting N₂ to . , the of to N₂ by under anaerobic conditions, had been observed earlier in the 1870s, but integration of these processes—fixation, nitrification, ammonification, and denitrification—into a cohesive biogeochemical cycle model solidified by the early 1900s, enabling predictive frameworks for soil fertility. Post-World War II advancements in stable isotope tracing, particularly with ¹⁵N, facilitated quantitative validation of global nitrogen fluxes starting in the 1940s-1950s. These techniques confirmed microbial discrimination in processes like denitrification, where lighter ¹⁴N is preferentially lost, enriching residual pools in ¹⁵N, and supported estimates of oceanic and terrestrial turnover rates. Recent analyses of Archean rocks, including 3.7 billion-year-old samples from the Isua Supracrustal Belt, reveal isotopic signatures indicating early aerobic nitrogen cycling, with evidence of nitrate production predating the Great Oxidation Event by up to 100 million years, refining models of primordial flux dynamics.

Core Biogeochemical Processes

Nitrogen Fixation

Nitrogen fixation is the biological or abiotic conversion of atmospheric dinitrogen (N₂), which constitutes about 78% of Earth's atmosphere, into reactive nitrogen compounds such as ammonia (NH₃) that organisms can assimilate. This process is essential because N₂'s triple bond renders it chemically inert and unavailable to most life forms without enzymatic or high-energy intervention. Biological nitrogen fixation, performed exclusively by prokaryotic diazotrophs, accounts for the majority of natural fixation, estimated at 100–200 Tg N year⁻¹ globally. These microorganisms employ the nitrogenase enzyme complex, comprising dinitrogenase (MoFe protein) and dinitrogenase reductase (Fe protein), which catalyzes the reduction: N₂ + 8 H⁺ + 8 e⁻ + 16 ATP → 2 NH₃ + H₂ + 16 ADP + 16 Pᵢ. The reaction demands a theoretical minimum of 16 ATP molecules per N₂ reduced, reflecting the high energy barrier of bond cleavage, which limits fixation rates in energy-constrained natural environments. Nitrogenase is highly sensitive to oxygen, which irreversibly inactivates its iron-sulfur clusters, necessitating anaerobic or microaerobic conditions; diazotrophs achieve this via spatial separation (e.g., heterocysts in cyanobacteria), temporal regulation, or protective mechanisms like high respiration rates. Diazotrophs include symbiotic species such as and , which form root nodules with leguminous plants, fixing up to kg N ha⁻¹ year⁻¹ in agricultural systems but contributing substantially to natural terrestrial inputs. Free-living diazotrophs like Azotobacter vinelandii and Clostridium species operate in soils and sediments without host symbiosis, while associative bacteria such as Azospirillum enhance fixation near plant roots. Globally, terrestrial biological fixation supplies approximately Tg N year⁻¹, with marine contributions from cyanobacteria like Trichodesmium adding another ~140 Tg N year⁻¹, though estimates vary due to methodological challenges in measuring rates. Abiotic fixation, though minor, occurs via lightning, which generates nitric oxide (NO) through high-temperature plasma arcs, subsequently oxidized to nitrate (NO₃⁻) and deposited as rain at 3–10 Tg N year⁻¹ worldwide. Other abiotic pathways, such as ultraviolet photolysis in the upper atmosphere or surface photochemical reactions on deserts, contribute negligibly compared to biological rates but are verifiable through nitrogen isotope ratios (¹⁵N/¹⁴N) enriched in fixed products relative to atmospheric N₂. These processes underscore fixation's role as the primary bottleneck in the nitrogen cycle, with biological mechanisms dominating due to their specificity and efficiency under ambient conditions.

Organic Nitrogen Transformations: Assimilation and Ammonification

Assimilation involves the incorporation of inorganic nitrogen, primarily ammonium (NH₄⁺) or nitrate (NO₃⁻), into organic forms such as amino acids by autotrophs like plants and heterotrophs including microbes. In plants and microorganisms, NH₄⁺ is assimilated via the glutamine synthetase (GS)/glutamate synthase (GOGAT) cycle, where GS catalyzes the formation of glutamine from NH₄⁺ and glutamate, and GOGAT recycles glutamate using 2-oxoglutarate. Nitrate assimilation requires prior reduction to nitrite (NO₂⁻) by nitrate reductase, followed by further reduction to NH₄⁺ by nitrite reductase, before entry into the GS/GOGAT pathway. This process is regulated by carbon availability and energy status, directly linking nitrogen uptake to photosynthetic carbon fixation and primary productivity, as assimilated nitrogen supports protein synthesis and growth. Ammonification, the mineralization of organic nitrogen to NH₄⁺, occurs during decomposition of biomass by heterotrophic bacteria and fungi, releasing nitrogen from compounds like proteins, nucleic acids, and urea through enzymatic hydrolysis and deamination. This step recycles internally cycled nitrogen, with global soil ammonification estimated to process substantial fluxes tied to turnover, often exceeding external inputs in natural systems. The process is strongly influenced by the carbon-to-nitrogen (C:N) ratio of substrates; microbial decomposition proceeds efficiently when C:N ratios are low (typically 10–25:1), allowing microbes to assimilate carbon for energy while mineralizing excess nitrogen as NH₄⁺, whereas high C:N ratios (>30:1) lead to nitrogen immobilization as microbes retain N for biomass. Carbon-nitrogen coupling is evident here, as decomposition drives both C mineralization to CO₂ and N release, with rates limited by microbial demand for balanced stoichiometry. Empirical from experiments highlight environmental controls on ammonification rates. Gross ammonification increases with , often doubling with each 10°C around 25–30°C, as enzymatic activity accelerates, though excessive (>40°C) inhibits microbes. positively affects rates by facilitating microbial activity and , with ammonification at 50–70% water-filled ; drier conditions (<30%) slow processes due to osmotic stress, while saturation promotes anaerobiosis and shifts to alternative pathways. These interactions underscore ammonification's sensitivity to microclimate, with combined temperature-moisture effects explaining seasonal variations in field measurements.

Nitrification

Nitrification is the aerobic, two-step microbial oxidation of ammonium (NH₄⁺) to nitrate (NO₃⁻), mediated by distinct groups of chemolithoautotrophic bacteria that derive energy from the oxidation of inorganic nitrogen compounds. In the first step, ammonia-oxidizing bacteria (AOB), such as species of Nitrosomonas, convert NH₄⁺ to nitrite (NO₂⁻) via the enzyme ammonia monooxygenase, yielding approximately 84 kcal/mol of free energy. The second step involves nitrite-oxidizing bacteria (NOB), such as Nitrobacter species, oxidizing NO₂⁻ to NO₃⁻ using nitrite oxidoreductase, which provides about 17 kcal/mol of energy—less efficient but essential for completing the process and preventing nitrite toxicity to other soil organisms. These bacteria are obligate aerobes, requiring molecular oxygen (O₂) as both an electron acceptor in respiration and a co-substrate in oxidation reactions; low oxygen levels below 0.5 mg/L significantly inhibit activity, confining nitrification primarily to well-oxygenated soils and surface waters. Optimal activity occurs at neutral to slightly alkaline pH (6.5–8.0), with rates declining sharply below pH 5 due to reduced enzyme functionality and bacterial viability, though acid-tolerant AOB strains exist in some environments. Synthetic inhibitors like nitrapyrin target ammonia monooxygenase, blocking the initial oxidation step and persisting in soils for 6–30 weeks depending on temperature and application timing, thereby reducing nitrate formation and associated losses. The overall reaction—NH₄⁺ + 2O₂ → NO₃⁻ + 2H⁺ + H₂O—releases protons (H⁺), contributing to net soil acidification; for every mole of ammonium oxidized, two equivalents of H⁺ are produced, exacerbating acidity in nitrogen-amended systems without base cation buffering. Metagenomic studies since 2015 have revealed complete ammonia-oxidizing (comammox) bacteria, primarily Nitrospira species, capable of performing both steps within a single organism, potentially simplifying the process in low-substrate environments and altering traditional views of nitrifier syntrophy. These discoveries highlight microbial diversity but do not negate the prevalence of canonical two-step pathways in most ecosystems.

Nitrogen Loss Processes: Denitrification, DNRA, and Anammox

Denitrification is a microbially mediated anaerobic process in which nitrate (NO₃⁻) is sequentially reduced to nitrite (NO₂⁻), nitric oxide (NO), nitrous oxide (N₂O), and ultimately dinitrogen gas (N₂), representing a primary pathway for fixed nitrogen loss from ecosystems to the atmosphere. This reduction is performed by facultative anaerobic heterotrophic bacteria, such as Pseudomonas and Paracoccus species, which utilize nitrate as an electron acceptor when oxygen is limited and organic carbon is available. The process predominates in waterlogged soils, sediments, and oxygen-deficient zones, with global terrestrial denitrification fluxes estimated at 100–250 Tg N yr⁻¹, accounting for approximately 50% of newly fixed nitrogen turnover on land. Incomplete denitrification often releases N₂O, a potent greenhouse gas with a 100-year global warming potential 273 times that of CO₂, contributing significantly to atmospheric N₂O concentrations. Dissimilatory nitrate reduction to ammonium (DNRA) provides an alternative anaerobic fate for nitrate, converting NO₃⁻ to ammonium (NH₄⁺) without nitrogen gas loss, thereby retaining fixed nitrogen within the ecosystem for potential reuse. This process is carried out by bacteria such as Sulfurimonas and certain Vibrio species under low-oxygen conditions with high organic carbon availability, where elevated carbon-to-nitrogen (C/N) ratios favor DNRA over denitrification by enhancing energy yields for the microbes involved. DNRA competes directly with denitrification for nitrate, with its prevalence increasing in carbon-rich environments like organic sediments or wastewater systems, though it typically contributes less to overall nitrogen dynamics compared to gas-producing pathways. In marine and freshwater sediments, DNRA can recycle up to 30–50% of nitrate under favorable conditions, influencing local nitrogen availability and microbial community structure. Anaerobic ammonium oxidation (anammox) is an autotrophic process where ammonium (NH₄⁺) and nitrite (NO₂⁻) are directly converted to N₂ gas, bypassing nitrate and representing a key nitrogen loss mechanism in anoxic environments. Performed by specialized planctomycete bacteria, such as Candidatus Scalindua in marine settings and Candidatus Brocadia in freshwater, anammox was first identified in wastewater systems in the early 1990s and later recognized as a major contributor to marine nitrogen removal. It accounts for substantial fixed nitrogen loss in ocean oxygen minimum zones (OMZs), contributing 30–50% of N₂ production in these regions and an estimated global marine flux of around 40–60 Tg N yr⁻¹. Anammox competes with heterotrophic denitrification and DNRA for nitrite substrates, with its efficiency enhanced in low-carbon, sulfide-limited anoxic niches, thereby shaping nitrogen budgets in stratified water columns and sediments. These loss processes collectively regulate ecosystem nitrogen levels, with their relative dominance determined by redox conditions, substrate availability, and microbial competition.

Ecosystem-Specific Variations

Terrestrial Nitrogen Dynamics

In terrestrial ecosystems, nitrogen dynamics hinge on intricate soil-plant-microbe interactions that facilitate fixation, assimilation, mineralization, and retention while minimizing losses. Biological nitrogen fixation dominates inputs, with symbiotic rhizobia in legume roots accounting for approximately 80% of biological N fixation in agricultural and natural systems, converting atmospheric N2 into ammonia via the nitrogenase enzyme. Free-living soil bacteria, such as Azotobacter and heterotrophic diazotrophs, contribute additional fixation without plant symbiosis, releasing fixed N into soil organic matter upon microbial death. These processes supply ammonium and organic N, which plants assimilate primarily through roots, often enhanced by mycorrhizal associations that extend hyphal networks for efficient uptake of inorganic and organic forms. Microbial decomposition of plant litter and root exudates drives ammonification, converting organic N back to ammonium, with rates modulated by litter quality—high C:N ratios (>25:1) promote microbial immobilization, retaining N in biomass, while low ratios accelerate mineralization. Arbuscular mycorrhizal fungi (AMF) further influence this by preferentially accessing ammonium from decomposing litter and inhibiting certain soil microbes, thereby boosting plant N acquisition efficiency and reducing mineralization losses. Nitrification follows, oxidizing ammonium to nitrate via soil bacteria like Nitrosomonas and Nitrobacter, though this step is oxygen-dependent and varies with soil aeration and pH. Terrestrial systems exhibit high N retention, with 93-97% of inputs sequestered in plant biomass, soil organic matter, and microbial pools in temperate forests, contrasting aquatic environments where dilution and export dominate. Leaching losses, primarily as nitrate, remain low at 1-5 kg N ha⁻¹ year⁻¹ in undisturbed temperate forests under ambient deposition but can rise to 35-40 kg N ha⁻¹ year⁻¹ in N-saturated or high-input sites, driven by exceedance of plant and microbial uptake capacity. Feedbacks from litter quality and mycorrhizae sustain this retention; for instance, ectomycorrhizal-dominated forests show slower decomposition of recalcitrant litter, enhancing long-term N immobilization compared to AMF systems. Process-based models like CENTURY simulate these terrestrial C-N interactions by partitioning soil organic matter into conceptual pools with distinct decomposition kinetics, incorporating microbial efficiency, lignin content, and N feedback on litter inputs to forecast dynamics over decades. The model has been validated for forest and grassland soils, revealing how coupled C-N cycles regulate N availability, with reduced litter quality under N limitation slowing turnover and amplifying retention. Such simulations underscore microbial primacy in terrestrial N cycling, where plant-microbe competition for ammonium shapes ecosystem responses to environmental shifts.

Freshwater Aquatic Systems

In freshwater aquatic systems, such as rivers and lakes, the nitrogen cycle processes terrestrial inputs through a combination of assimilation by primary producers, microbial transformations, and losses via denitrification and burial in sediments. Nitrogen enters these systems primarily as nitrate and ammonium from upstream watersheds, where it supports phytoplankton growth before being remineralized or converted. Sediments serve as hotspots for denitrification, where anaerobic bacteria reduce nitrate to dinitrogen gas, removing an estimated 20% of global denitrification occurs in freshwater systems, effectively mitigating a substantial portion of land-derived nitrogen before it reaches estuaries. This removal efficiency varies by hydrology, with slower-flowing rivers and stratified lakes exhibiting higher rates due to prolonged contact with sediments. Recent assessments highlight the disproportionate of inland waters in , contributing approximately 10% of naturally fixed despite occupying less than 1% of Earth's surface area. Biological fixation, primarily by diazotrophic cyanobacteria during blooms in nutrient-enriched lakes and reservoirs, accounts for this outsized input, with rates amplified in phosphorus-limited environments where scarcity favors N2-fixing . A 2025 estimates that inland systems alone fix tens of teragrams of annually, underscoring their underappreciated on downstream and . These blooms often coincide with eutrophication hotspots triggered by episodic runoff pulses, which elevate reactive and , fostering algal and subsequent accumulation. Seasonal stratification in lakes drives hypoxia cycles that modulate nitrogen processing, as warmer surface waters inhibit vertical mixing, leading to oxygen depletion in bottom layers from organic decomposition. Under hypoxic conditions, denitrification and anaerobic ammonium oxidation intensify, enhancing nitrogen loss, while prolonged stratification—exacerbated by warming trends—extends anoxic periods and alters microbial community structure. In temperate lakes, this stratification typically peaks in summer, creating redox gradients that favor dissimilatory nitrate reduction over aerobic nitrification, thereby influencing the balance between nitrogen retention and export. Such dynamics contribute to variable nitrogen budgets, with shallower systems experiencing more frequent turnover events that reoxygenate sediments and reset transformation pathways.

Marine Nitrogen Cycle

The is characterized by processes in the and sediments, driven by vertical mixing, , and oxygen gradients to saline environments. by relies on , distinguished as new production supported by external inputs like dinitrogen (N2) fixation and from waters, versus regenerated production from recycled within the euphotic . This , formalized in the , underscores that new fuels carbon export to depth, while regenerated production sustains , with the f- (new to ) varying from near in oligotrophic gyres to higher values in regions. Vertical mixing and eddy diffusion supply regenerated to surface waters, balancing losses in oxygen minimum zones (OMZs). Marine N2 fixation, primarily by diazotrophic cyanobacteria such as Trichodesmium and UCYN-A, inputs an estimated 100–190 Tg N per year globally, compensating for denitrification and anaerobic ammonium oxidation (anammox) losses to maintain steady-state nitrogen inventories. Recent syntheses refine open-ocean rates to around 140 Tg N year−1, with coastal and inland waters contributing an additional ~40 Tg N year−1, though methodological artifacts in tracer assays have led to debates over coastal overestimation in prior models. In sediments, dissimilatory nitrate reduction and coupled nitrification-denitrification dominate under diffusive oxygen penetration, distinct from water-column dynamics due to salinity-influenced microbial consortia. In OMZs, such as those in the eastern tropical Pacific and Arabian Sea, anammox drives 30–50% of total nitrogen loss, oxidizing ammonium with nitrite to N2 without nitrate production, often exceeding canonical denitrification. This process, mediated by Candidatus Scalindua bacteria, relies on proximate nitrification sources and accounts for up to 40% of oceanic fixed nitrogen removal. Ocean stratification and warming may expand OMZs, amplifying losses, while upwelling regenerates nitrate for surface productivity in productive margins. Ocean acidification, via lowered pH from CO2 uptake, inhibits nitrification by ammonia-oxidizing bacteria and archaea, potentially reducing rates by 3–44% over decades due to proton sensitivity in ammonia uptake. This feedback could diminish nitrate supply for new production, altering N2O emissions and exacerbating nitrogen limitation in surface waters, though compensatory shifts in diazotrophy remain uncertain. Empirical mesocosm studies confirm pH thresholds below 7.8 slowing oxidation steps, with implications for coupled redox cascades.

Global Scale and Natural Fluxes

Overall Nitrogen Budget and Reservoirs

The dominant nitrogen reservoir on Earth is the atmosphere, containing approximately $4 \times 10^{21} grams of dinitrogen (N_2), which constitutes about 78% of the atmospheric mass. In comparison, the global pool of reactive nitrogen (Nr)—including species like ammonia (NH_3), nitrogen oxides (NO_x), nitrates (NO_3^-), and organic forms—is orders of magnitude smaller, estimated at around $3.9 \times 10^{12} grams, distributed primarily across soils (roughly 10,000–20,000 Tg N), ocean dissolved inorganic N (~600 Tg), and biomass. These Nr reservoirs represent the bioavailable fraction that cycles through ecosystems, contrasting sharply with the inert atmospheric N_2 pool. Pre-industrial global Nr creation totaled approximately 100–200 Tg N yr^{-1}, driven mainly by natural biological nitrogen fixation (58–100 Tg on land and ~140 Tg in oceans), lightning (~5–20 Tg), and minor abiotic sources, with losses balanced by denitrification and sedimentation. In the modern era, total Nr inputs have nearly doubled to ~350–400 Tg N yr^{-1}, with anthropogenic contributions reaching 226 Tg N yr^{-1} by 2020, primarily from the Haber-Bosch process (~120 Tg), crop-associated fixation (~30 Tg), and fossil fuel combustion (~20–46 Tg NO_x). Natural fluxes continue to turnover 200–300 Tg N yr^{-1} through fixation and subsequent transformations, but human perturbations have disrupted the balance, leading to Nr accumulation in environmental compartments. A key long-term sink in the global nitrogen budget is burial in sediments, estimated at ~200 Tg N yr^{-1} across marine and terrestrial systems, where organic and inorganic N is sequestered beyond biological recycling. Isotopic analysis of ^{15}N/^{14}N ratios in sediments and archives provides evidence for these pathways, distinguishing denitrified N losses (enriched in ^{15}N) from fixed inputs and revealing imbalances, such as elevated modern burial rates in lakes (~9.6 Tg N yr^{-1} total, with ~7 Tg anthropogenic). This sedimentary flux, alongside gaseous losses, maintains approximate steady-state on geological timescales but highlights amplified Nr retention under current human-influenced budgets.

Microbial and Abiotic Drivers

Microbial communities drive the core transformations of the nitrogen cycle through specialized functional guilds, with nitrogen fixation primarily mediated by diazotrophs harboring the nifH gene, which encodes the Fe protein subunit of nitrogenase. Metagenomic surveys of global soils have identified diverse nifH sequences across bacterial and archaeal phyla, revealing that non-cyanobacterial diazotrophs predominate in terrestrial environments, contributing substantially to fixed nitrogen inputs. Similarly, denitrification relies on nosZ genes encoding nitrous oxide reductase, where atypical nosZ variants—often associated with uncultured taxa—outnumber typical forms in soils, enabling N2O reduction under varied conditions and potentially accounting for over half of denitrifying potential in anaerobic microsites. Recent genome-resolved metagenomics from oxygen-deficient zones and rhizospheres has uncovered uncultured microbial lineages, such as novel metagenome-assembled genomes (MAGs) from Planctomycetes and other rare phyla, that perform key steps like partial denitrification and anammox, driving more than 50% of nitrogen turnover in stratified ecosystems where culturable organisms are underrepresented. Abiotic factors impose physicochemical constraints on these microbial processes, with soil pH strongly influencing guild activity: nitrification optima occur near neutral pH (6.5-7.5), while acidic conditions (pH <5) favor denitrification over ammonia oxidation, altering net nitrogen retention. Redox potential dictates process succession, as aerobic environments (Eh >300 mV) support nitrification by ammonia-oxidizing bacteria, whereas suboxic to anoxic conditions (Eh <200 mV) activate denitrifiers and inhibit oxygen-sensitive nitrogenase, creating spatial heterogeneity in soils and sediments. Temperature modulates enzyme kinetics and microbial growth, with denitrification rates peaking at 25-35°C due to enhanced reductase activity, though exceeding 40°C inhibits most mesophilic guilds, reducing overall cycle efficiency in warming scenarios. Lightning-induced fixation represents a minor abiotic input, estimated at less than 5 Tg N per year globally through NOx formation and deposition, verifiable via isotopic signatures but negligible compared to biological fixation. Stoichiometric imbalances, particularly deviations from the Redfield ratio of 16:1 (N:P atomic), constrain nitrogen transformation rates by limiting co-substrate availability; phosphorus scarcity relative to nitrogen promotes fixation to restore balance in phosphorus-replete systems, while excess nitrogen relative to phosphorus accelerates losses via denitrification, enforcing causal limits on cycle throughput independent of microbial abundance. These drivers interact synergistically, as metagenomic evidence indicates that uncultured taxa exhibit broader abiotic tolerances, sustaining transformations under suboptimal pH or redox gradients where canonical microbes falter.

Natural Variability and Feedbacks

Climate-driven variations significantly influence nitrogen cycle processes, particularly through temperature and precipitation effects on mineralization and fixation rates. In grasslands, net soil nitrogen mineralization increases with rising temperatures, with studies showing positive correlations where rates can rise by approximately 5-10% per 1°C under optimal moisture conditions, reflecting enhanced microbial decomposition of organic matter. Increased precipitation generally accelerates mineralization by improving soil moisture for microbial activity, but extreme droughts suppress biological nitrogen fixation, with reductions up to 88.8% observed under severe water deficits in legume systems due to impaired diazotroph activity and nodule function. Negative feedbacks stabilize nitrogen availability, preventing excessive accumulation in ecosystems. In regions approaching natural nitrogen saturation, primary production often shifts from nitrogen limitation to phosphorus limitation, as evidenced by long-term ecosystem development models where phosphorus weathering rates determine transition timelines, thereby constraining further nitrogen fixation demands. Evolutionary adaptations, such as the development of root nodule symbiosis in legumes with rhizobia, enhance nitrogen fixation efficiency in nitrogen-poor soils, with horizontal gene transfer enabling this trait's spread and providing a selective advantage that buffers against variability in abiotic nitrogen inputs. Over glacial-interglacial cycles, nitrogen deposition exhibits marked fluctuations, with ice core records from Antarctica revealing higher δ¹⁵N values in nitrate during glacial periods (differing by ~26.6‰ from Holocene levels), indicative of altered atmospheric processing and increased aeolian dust transport contributing to terrestrial nitrogen inputs. These variations, driven by changes in global dust flux and ocean-atmosphere exchange, underscore geological timescales of feedback where enhanced glacial dust deposition temporarily elevated bioavailable nitrogen before interglacial reductions.

Human Interventions in the Cycle

Industrial Nitrogen Fixation via Haber-Bosch Process

The Haber-Bosch process catalyzes the synthesis of ammonia (NH₃) from dinitrogen (N₂) and dihydrogen (H₂) via the reversible reaction N₂ + 3H₂ ⇌ 2NH₃, which is exothermic (ΔH = -92 kJ/mol) and decreases the number of gas moles, favoring product formation at high pressures and low temperatures per Le Chatelier's principle. Industrial conditions compromise between thermodynamics and kinetics, operating at 400–500°C and 150–300 bar with a heterogeneous iron-based catalyst (typically Fe₂O₃ promoted by K₂O, Al₂O₃, and CaO) to achieve single-pass conversions of 10–15%, recycled for overall yields exceeding 95%. This technological intervention circumvents the kinetic inertness of the N≡N triple bond (bond energy 941 kJ/mol), enabling fixation rates orders of magnitude beyond symbiotic or free-living microbial processes constrained by enzyme efficiencies and environmental factors. Developed by Fritz Haber (lab-scale, 1909) and scaled industrially by Carl Bosch (1913) at BASF, the process marked a pivotal decoupling of nitrogen availability from natural biogeochemical limits, transforming inert atmospheric N₂ (78% of air) into reactive forms on demand. By 2020, global ammonia production via Haber-Bosch fixed approximately 150 Tg N annually, primarily as NH₃ for downstream conversion to fertilizers like urea and ammonium nitrate. This output, equivalent to or exceeding natural biological fixation fluxes, underpinned exponential crop yield increases from the 1910s onward; without it, global grain production would have stagnated at early 20th-century levels, limiting food supply for billions. The process demands substantial energy input, accounting for 1–2% of global primary energy consumption, with hydrogen sourced mainly from natural gas steam methane reforming (endothermic, requiring ~30 GJ/tonne NH₃) and synthesis gas compression/heating adding further demands. Efforts to mitigate fossil fuel dependence include "green" variants integrating electrolytic H₂ from water electrolysis powered by renewables, but these comprised less than 0.1% of production in the early 2020s due to electrolysis costs exceeding $3/kg H₂ and scalability challenges. Pilot plants, such as those operational by 2023 producing ~185,000 tonnes NH₃/year electrolytically, demonstrate feasibility but highlight persistent economic barriers relative to conventional efficiencies. Ongoing catalyst research targets lower-temperature operation to reduce energy penalties, yet core thermodynamics necessitate high-pressure containment, limiting near-term disruptions to the established paradigm.

Agricultural Fertilization and Crop Management

Nitrogen fertilizers commonly applied in agriculture include urea, ammonium nitrate, and ammonium-based compounds, which provide nitrogen in forms readily convertible to plant-available ammonium or nitrate through soil processes. Global consumption of synthetic nitrogen fertilizers for croplands reaches approximately 100 Tg N per year, with projections indicating continued demand growth to support expanding food production. Nitrogen use efficiency (NUE) in cereal crops, defined as the percentage of applied nitrogen recovered in grain yield, averages 33-40% worldwide, meaning a substantial portion remains unaccounted for due to losses including volatilization of ammonia, which can account for 10-20% of applied nitrogen depending on soil pH, moisture, and application method. Strategies to mitigate volatilization involve incorporating fertilizers into soil or using urease inhibitors, enhancing overall agronomic efficiency. During the Green Revolution from the 1960s to 1980s, the integration of high-yielding hybrid cereal varieties responsive to nitrogen inputs with increased fertilizer application doubled or tripled grain yields in regions like South Asia and Mexico, enabling support for billions in population growth without proportional land expansion. Contemporary crop management employs site-specific practices such as variable-rate nitrogen application, guided by soil testing, remote sensing, and crop models, which can improve nitrogen uptake by 10-20% through tailored dosing that matches spatial variability in soil fertility and crop needs. Emerging microbiome engineering trials in the 2020s aim to enhance biological nitrogen fixation in non-legume cereals by inoculating roots with engineered symbiotic bacteria, potentially reducing reliance on synthetic inputs while maintaining yields.

Other Anthropogenic Sources: Combustion and Waste

Anthropogenic nitrogen emissions from fossil fuel combustion primarily occur as nitrogen oxides (NOx), formed via the thermal fixation of atmospheric N₂ at high temperatures exceeding 1,300°C in engines, boilers, and industrial processes. Global anthropogenic NOx emissions, largely attributable to fossil fuel combustion in transportation, power generation, and manufacturing, totaled approximately 46.5 Tg N yr⁻¹ (with uncertainty of ±2.7 Tg N yr⁻¹) during 2016–2020. These fluxes stem from oxidation reactions like the Zeldovich mechanism, where N₂ + O → NO followed by NO + ½O₂ → NO₂, yielding NO and NO₂ in roughly equal proportions under typical combustion conditions. Relative to agricultural sources, combustion-derived NOx constitutes a minor share of total anthropogenic reactive nitrogen creation, which surpasses 170 Tg N yr⁻¹ globally, with fertilizers dominating at over 100 Tg N yr⁻¹ applied annually. Wastewater and sewage handling introduce reactive nitrogen mainly as ammonium (NH₄⁺) from human excreta and organic waste, with global human nitrogen excretion estimated at 23 Tg N yr⁻¹. Untreated or inadequately processed effluents discharge this nitrogen into waterways, though advanced treatment mitigates releases; for instance, activated sludge systems achieve 50–90% nitrogen removal via aerobic nitrification (NH₄⁺ to NO₃⁻ by autotrophic bacteria like Nitrosomonas and Nitrobacter) followed by anoxic denitrification (NO₃⁻ to N₂ by heterotrophs like Pseudomonas). Net emissions from wastewater remain secondary to combustion NOx in scale, estimated at around 10–15 Tg N yr⁻¹ entering environments after partial treatment globally, far below fertilizer-driven inputs. Historical peaks in NOx and ammonia emissions from these sources contributed to acid rain deposition, with sulfate and nitrate wet deposition rates in eastern North America reaching 20–30 kg N ha⁻¹ yr⁻¹ in the 1980s before regulatory interventions like the U.S. Clean Air Act Amendments of 1990 reduced emissions by over 50%. Similar declines, tracked by European Monitoring and Evaluation Programme networks, reflect scrubber technologies and fuel switching, lowering deposition to below 10 kg N ha⁻¹ yr⁻¹ in many regulated areas by the 2010s.

Balanced Assessment of Human Impacts

Achievements: Enhanced Food Security and Economic Growth

Synthetic nitrogen fertilizers have substantially increased global crop yields, supporting food production sufficient to feed nearly half of the world's population. Analyses indicate that nitrogen from synthetic sources accounts for approximately 48% of food production as of the late 2000s, enabling the sustenance of billions amid population growth from 2.5 billion in 1950 to over 8 billion today. Global cereal output rose from 877 million metric tons in 1961 to 2,794 million metric tons in 2022, with synthetic N applications expanding tenfold since 1960 to drive this yield intensification on existing arable land. Without such inputs, counterfactual models project cereal yields 30-50% lower, as natural nitrogen fixation alone cannot match demand for high-yield monocultures. In regions prone to food scarcity, nitrogen additions have directly averted famines by transforming low-productivity agriculture. In India, rice yields increased from around 1.0-1.3 metric tons per hectare in the early 1960s to over 2.5 metric tons per hectare by the 2010s, propelled by the Green Revolution's integration of synthetic N with high-yielding varieties, shifting the country from chronic imports and famine risks to self-sufficiency and net exporter status. This yield surge, crediting fertilizers for over 50% of gains in staple crops, prevented mass starvation forecasted in the 1960s amid rapid population growth. Economically, nitrogen investments deliver high returns, particularly in developing nations where agriculture employs much of the workforce. For each dollar spent on N fertilizers, returns range from $4 to $6 in added crop value, boosting farmer incomes, rural economies, and poverty alleviation by enhancing productivity on smallholder farms. These gains have underpinned broader growth, with fertilizer-enabled agriculture contributing to GDP rises exceeding 6% annually in the 1960s across expanding economies, while sustaining food security for urbanizing populations.

Environmental Costs: Eutrophication, Soil Degradation, and Biodiversity Loss

Excess reactive nitrogen from agricultural runoff and deposition drives eutrophication in freshwater and coastal systems by fueling excessive algal growth, which upon decay consumes dissolved oxygen and creates hypoxic zones. In the Gulf of Mexico, this manifests as an annual "dead zone" spanning approximately 13,000 to 22,000 km² in recent measurements, primarily resulting from Mississippi River Basin inputs exceeding natural nutrient fluxes, with nitrogen loads from fertilizers contributing over 70% of the flux. Hypoxia levels below 2 mg/L O₂ displace mobile species like shrimp and fish while stressing benthic organisms, with recovery lagging behind nutrient reductions due to sediment legacy effects. Critical nitrogen application thresholds for avoiding such coastal eutrophication are estimated at 10-20 kg N ha⁻¹ yr⁻¹ in vulnerable watersheds, beyond which algal biomass surges and oxygen minima intensify. Nitrogen excess degrades soils through acidification and leaching, as ammonium-based fertilizers undergo nitrification (NH₄⁺ → NO₃⁻ + 2H⁺), releasing protons that lower pH by 0.2-0.5 units per decade in high-input croplands (>100 kg N ha⁻¹ yr⁻¹). This proton accumulation depletes base cations like Ca²⁺ and Mg²⁺, elevating exchangeable Al³⁺ to toxic levels (>1 cmol₊ kg⁻¹ soil), which inhibits root growth and microbial activity essential for nutrient cycling. Concurrently, nitrate leaching from unsaturated zones transports 20-50 kg N ha⁻¹ yr⁻¹ in fertilized systems, raising shallow groundwater NO₃⁻-N concentrations above 10 mg L⁻¹—far exceeding natural baselines (<3 mg L⁻¹)—and risking denitrification hotspots that emit N₂O. Recovery requires decades of reduced inputs and liming, as observed in European trials where pH rebounded only 0.1-0.3 units after 10-20 years of N withholding. Elevated nitrogen deposition disrupts terrestrial biodiversity by favoring competitive, nitrophilous species (e.g., grasses like Urtica dioica) over slow-growing natives adapted to low-nutrient conditions, leading to community homogenization. In European grasslands, deposition rates >10 kg N ha⁻¹ yr⁻¹ correlate with 15-25% declines in species richness over decades, as documented in resurveyed plots from the UK and Netherlands, where oligotrophic forbs like Succisa pratensis vanish while total cover shifts to eutrophic dominants. Thresholds for initial diversity loss range 5-15 kg N ha⁻¹ yr⁻¹ across alliances, with cascading effects reducing pollinator support and soil fauna via altered litter quality. Long-term exclusion experiments confirm partial reversibility, but legacy soil N pools delay native recovery by 10-20 years post-reduction.

Atmospheric and Climatic Interactions

Nitrous oxide (N₂O), produced primarily through microbial processes in nitrogen-enriched soils, accounts for approximately 6% of total anthropogenic radiative forcing from greenhouse gases, with a global warming potential 265–298 times that of CO₂ over 100 years. Agricultural soils contribute about 52–56% of global anthropogenic N₂O emissions, largely from denitrification and nitrification following fertilizer application and manure management. With an atmospheric lifetime of 114–116 years, N₂O accumulates persistently, amplifying long-term warming; its emissions rose 40% from 1980 to 2020, reaching 18.5 Tg N yr⁻¹ globally. Nitrogen oxides (NOx), emitted from combustion and soil processes, drive tropospheric ozone formation, a short-lived climate forcer contributing to radiative forcing and surface smog. Ozone's greenhouse effect stems from NOx-initiated photochemical reactions with volatile organic compounds, enhancing warming in polluted regions; NOx reductions can paradoxically increase ozone in VOC-limited areas due to altered chemistry. Globally, anthropogenic NOx emissions averaged 46.5 Tg N yr⁻¹ during 2016–2020, predominantly from fossil fuel combustion. Reactive nitrogen deposition to land totaled 92.7 Tg N yr⁻¹ in 2020, delivered via wet (e.g., ammonium nitrate in rainfall) and dry pathways, with oxidized forms (NOy) declining post-2016 due to emission controls while reduced forms (NHx) rose in some regions. This deposition acidifies forest soils and precipitation by contributing nitrate and ammonium ions, lowering pH and mobilizing aluminum, which stresses tree roots and alters microbial communities. Short-term fertilization effects, however, enhance net primary productivity in nitrogen-limited forests, boosting carbon sequestration by 2–5% in some ecosystems before saturation leads to diminished returns or losses via N₂O feedbacks. Global models indicate total deposition peaked around 2000 and has since stabilized or regionally declined, mitigating some acidification but sustaining elevated baseline fluxes that interact with warming to reduce soil carbon stability.

Direct Effects on Human Health

High concentrations of nitrate in drinking water pose a risk of methemoglobinemia, a condition impairing hemoglobin's oxygen transport, particularly affecting infants under six months and manifesting as cyanosis or "blue baby syndrome." The U.S. Environmental Protection Agency has established a maximum contaminant level of 10 mg/L for nitrate-nitrogen (NO3-N) to mitigate this acute effect. Epidemiological data indicate such cases are exceedingly rare, with fewer than 0.1% attributable primarily to water sources; gastrointestinal infections represent the dominant causal factor, rendering the incremental risk from nitrate exposure negligible in most assessments. Links between dietary or waterborne nitrates and cancer involve endogenous formation of N-nitrosamines, potent carcinogens generated via nitrosation of amines in the stomach under acidic conditions. The International Agency for Research on Cancer classifies ingested nitrate and nitrite as probably carcinogenic to humans (Group 2A) when resulting in such nitrosation, based on animal bioassays showing tumors in multiple organs and limited human evidence for gastric and esophageal cancers. However, direct causation remains contested, as confounding variables like dietary patterns, Helicobacter pylori infection, and antioxidant intake (e.g., vitamin C inhibiting nitrosation) complicate observational studies; meta-analyses often fail to establish consistent dose-response relationships beyond high-exposure cohorts. Nitrogen oxides (NOx), emitted from combustion processes altering the nitrogen cycle, contribute to ground-level ozone and secondary particulate matter (PM2.5) formation, triggering oxidative stress in lung tissues and exacerbating conditions like asthma, bronchitis, and reduced lung function. Short-term NOx exposure correlates with increased hospital admissions for respiratory distress, while chronic exposure elevates risks for preterm birth and developmental impairments in children. Globally, the World Health Organization attributes approximately 4.2 million premature deaths in 2019 to ambient air pollution, with NOx as a key precursor to ozone and PM implicated in roughly 15-20% of these cardiopulmonary outcomes based on integrated exposure-response models. Direct health risks from reactive nitrogen are thus pathway-specific and dose-dependent, with waterborne nitrates primarily acute but infrequent, and airborne NOx chronically insidious via secondary pollutants; these must be weighed against indirect benefits from nitrogen-enhanced agriculture, which has averted widespread protein-energy malnutrition by tripling global crop yields since 1960, yielding a net positive caloric security despite localized exposures.

References

  1. [1]
    The Nitrogen Cycle: A Large, Fast, and Mystifying Cycle - PMC - NIH
    Sep 25, 2019 · The nitrogen cycle is essential for life, involving nitrogen fixation and denitrification, and has two main pools: atmospheric and reactive  ...
  2. [2]
  3. [3]
    The Nitrogen Cycle | Earth Science - Visionlearning
    The five processes in the nitrogen cycle – fixation, uptake, mineralization, nitrification, and denitrification – are all driven by microorganisms. Humans ...Nitrogen Fixation · Nitrogen Mineralization · Denitrification
  4. [4]
    Nitrogen | N (Element) - PubChem
    Nitrogen was discovered by chemist and physician Daniel Rutherford in 1772. He removed oxygen and carbon dioxide from air and showed that the residual gas would ...
  5. [5]
    [PDF] Daniel Rutherford, Nitrogen, - UNT Digital Library
    Apr 6, 2016 · as nitrogen. The characterization of "malignant air." Daniel Rutherford described the discovery of this new air in his 1772 M.D. dissertation.
  6. [6]
    Nitrogen | Podcast - Chemistry World
    Cavendish prepared nitrogen gas by this means. He passed air back and forth over heated charcoal which converted the oxygen in the air to carbon dioxide. The ...
  7. [7]
    Nitrogen or Azote Facts - ThoughtCo
    Jan 31, 2019 · The French chemist Antoine Laurent Lavoisier named nitrogen azote, meaning without life. Properties: Nitrogen gas is colorless, odorless, and ...
  8. [8]
    Nitrogen - Element information, properties and uses | Periodic Table
    However, in his list of the then known elements, Lavoisier included the term azote or azotic gas for what we now call nitrogen. This again stems from Greek ...
  9. [9]
    A chronology of human understanding of the nitrogen cycle - Journals
    Jul 5, 2013 · Henry Cavendish produced HNO3 in 1785 by passing sparks through a jar of air confined over water. Between 1772 and 1780, Joseph Priestley ...
  10. [10]
    Law of the Minimum - Soil and Environmental Sciences
    Mar 6, 2000 · Justus von Liebig's Law of the Minimum states that yield is proportional to the amount of the most limiting nutrient, whichever nutrient it may be.
  11. [11]
    The Law of the Minimum | Mosaic Crop Nutrition
    The Law of the Minimum, which states that if one of the essential plant nutrients is deficient, plant growth will be poor even when all other essential ...
  12. [12]
    Liebig Advocates Artificial Fertilizers | Research Starters - EBSCO
    In his influential 1840 book, Liebig emphasized the importance of minerals—like potassium and phosphorus—along with nitrogen for optimal plant growth. He argued ...Key Figures · Summary Of Event · Significance
  13. [13]
    Contribution to the HiStory series in Plant Nutrition
    Mar 16, 2025 · Lawes and Gilbert carried out numerous field trials that proved that nitrogen fertilization significantly increased the yields of many crops, ...<|separator|>
  14. [14]
    Lawes and Gilbert: an unlikely Victorian agricultural partnership
    Jan 9, 2018 · Lawes & Gilbert demonstrated that they acquired nitrogen from the soil, whereas Justus von Liebig, in Germany, argued that they obtained it from ...Missing: fertility 1850s
  15. [15]
    Biological nitrogen fixation: rates, patterns and ecological controls in ...
    Examining all of these sensitivities, we suggest that a reasonable range for pre-industrial terrestrial BNF is 40–100 Tg N yr−1, with a preferred single ...
  16. [16]
    Biological nitrogen fixation: rates, patterns and ecological controls in ...
    Jul 5, 2013 · Our estimate is that pre-industrial N fixation was 58 (range of 40 ... N reduces our best estimate to 44 Tg N yr−1. This approach ...
  17. [17]
    Capitalism 'Solves' the Nitrogen Crisis: A Brief History
    Sep 30, 2019 · Capitalism, focused on profit, undermined traditional methods, shifted food production, and used guano and sodium nitrate to address the  ...Missing: scarcity pre-
  18. [18]
    Haber-Bosch process | Definition, Conditions, Importance, & Facts
    Sep 26, 2025 · Haber-Bosch process, method of directly synthesizing ammonia from hydrogen and nitrogen, developed by the German physical chemist Fritz Haber.
  19. [19]
    [PDF] The Haber-Bosch Heritage: The Ammonia Production Technology
    The initial operation of the commercial plant commis- sioned in September 1913 was based on hydrogen and nitrogen produced by this cryogenic separation, but.
  20. [20]
    [PDF] How a century of ammonia synthesis changed the world
    Sep 28, 2008 · 100 Tg N from the Haber–Bosch process was used in global agriculture, whereas only 17 Tg N was consumed by humans in crop, dairy and meat.
  21. [21]
    Sergei Winogradsky: a founder of modern microbiology and the first ...
    He isolated the first pure cultures of the nitrifying bacteria and confirmed that they carried out the separate steps of the conversion of ammonia to nitrite ...Abstract · Winogradsky's early years · Zurich and nitrification · Back to Russia
  22. [22]
    Martinus W. Beijerinck | Biography, Virology, & Facts - Britannica
    Oct 13, 2025 · He also discovered new types of bacteria from soil and described biological nitrogen fixation (the conversion of nitrogen gas into ammonium ...Missing: cycle conceptualization
  23. [23]
    A chronology of human understanding of the nitrogen cycle - PMC
    In short order, the processes of nitrification, biological nitrogen fixation (BNF) and denitrification were discovered. The process of nitrification was ...
  24. [24]
    [PDF] Isotopic Tracers of the Marine Nitrogen Cycle: Present and Past
    Unlike N2 fixation, denitrification strongly discriminates against 15N (large isotope fractionation potential). ... made to quantify global ocean N fluxes and ...
  25. [25]
    3Q's: Demystifying Earth's Ancient Nitrogen Cycle - MIT EAPS
    May 16, 2025 · Samples from some of Earth's oldest rocks show evidence of an aerobic nitrogen cycle 100 million years earlier than previously thought.
  26. [26]
    Reconstructing Nitrogen Sources to Earth's Earliest Biosphere at 3.7 ...
    Here we revisit the Isua meta-turbidites and provide new data for metal abundances as well as organic carbon and nitrogen isotope values.<|control11|><|separator|>
  27. [27]
    An Earth-system perspective of the global nitrogen cycle - Nature
    Jan 16, 2008 · On a global basis, this is more than that supplied by natural biological nitrogen fixation on land (110 Tg N per year) or in the ocean (140 Tg N ...
  28. [28]
    The Abiotic Nitrogen Cycle | ACS Earth and Space Chemistry
    Aug 16, 2017 · These processes collectively define a natural chemical nitrogen cycle, analogous to the familiar biologically driven cycle but even more ...Abstract · Keywords · Figure 1<|separator|>
  29. [29]
    Evidence of limited N2 fixation in the Southern Ocean - Nature
    Apr 5, 2025 · However, estimates of global N2 fixation harbor substantial uncertainties, ranging from less than 100 to over 200 Tg N year−12,3,4.
  30. [30]
    Nitrogenase - an overview | ScienceDirect Topics
    The theoretical cost for reducing one molecule of N2 with the concomitant reduction of 2H+ is 16ATP. However, under natural conditions the cost may be even ...<|separator|>
  31. [31]
    5.15D: Anaerobiosis and N₂ Fixation - Biology LibreTexts
    Nov 23, 2024 · The iron (Fe) found in nitrogenases is very sensitive to oxygen, if there is too much oxygen this will in the end disrupt nitrogenase function.
  32. [32]
    Nitrogen-Fixing Bacterium - an overview | ScienceDirect Topics
    Nitrogen-fixing rhizobacteria fix nitrogen symbiotically (Rhizobium) and nonsymbiotically (Azotobacter, Azospirillum, Beijerinckia, etc.). Symbiotic nitrogen ...
  33. [33]
    Global Nitrogen: Cycling out of Control - PMC - NIH
    Lightning accounts for some naturally occurring reactive nitrogen—worldwide each year, lightning fixes an estimated 3–10 teragrams (Tg), the usual measurement ...Missing: abiotic | Show results with:abiotic
  34. [34]
    The Role of Glutamine Synthetase (GS) and Glutamate Synthase ...
    The glutamine synthetase (GS)/glutamate synthase (GOGAT) cycle represents a crucial metabolic step of N assimilation, regulating crop yield.
  35. [35]
    Nitrogen assimilation in plants: current status and future prospects
    Ammonium assimilation is catalyzed by the glutamine synthetase (GS)/glutamine-2-oxoglutarate aminotransferase (GOGAT, also known as glutamate synthase) cycle, ...
  36. [36]
    Genomic organization and expression profiles of nitrogen ...
    Jun 24, 2024 · Within plants, nitrogen is further incorporated into amino acids through the glutamine synthetase/glutamate synthase (GS/GOGAT) cycle, which is ...
  37. [37]
    The importance of cytosolic glutamine synthetase in nitrogen ...
    Apr 16, 2009 · Glutamine synthetase assimilates ammonium into amino acids, thus it is a key enzyme for nitrogen metabolism. The cytosolic isoenzymes of ...
  38. [38]
    Ammonification - an overview | ScienceDirect Topics
    This process is affected by temperature, pH, C/N ratio, nutrient content, and soil conditions. Optimum pH area is between 6.5 and 8.5 and temperature ...
  39. [39]
    A global dataset of gross nitrogen transformation rates across ...
    Sep 19, 2024 · We compiled a dataset of gross nitrogen transformation rates (GNTR) of mineralization, nitrification, ammonium immobilization, nitrate immobilization, and ...
  40. [40]
    Ammonification - an overview | ScienceDirect Topics
    So, the process of ammonification is a rate-limiting step in the nitrogen cycle. ... An optimum carbon-to-nitrogen (C/N) ratio helps in bacterial growth ...Ii Reservoirs And Fluxes Of... · The Oceans And Marine... · 8.11. 2.1 Biogeochemical...
  41. [41]
    The Significance of Microbial Transformation of Nitrogen ... - MDPI
    Research has shown that the following environmental factors influence the ammonification process: temperature, pH, C:N ratio, the content of nutrients and soil ...4.2. Nitrification · 4.3. Denitrification · 4.5. Anammox
  42. [42]
    Effects of Temperature and Humidity on Soil Gross Nitrogen ...
    Feb 25, 2023 · Both temperature and humidity significantly affected the N transformations. The gross ammonification and nitrification were significantly ...Missing: empirical | Show results with:empirical
  43. [43]
    Temperature and substrate effects on C & N mineralization and ...
    At high temperatures, ammonification rates increased in the shallow soil but decreased in the deep soil. The increase in the incubation temperature resulted ...
  44. [44]
    Effects of Temperature and Humidity on Soil Gross Nitrogen ...
    Feb 25, 2023 · Both temperature and humidity significantly affected the N transformations. The gross ammonification and nitrification were significantly ...Missing: empirical | Show results with:empirical
  45. [45]
    The Nitrogen Cycle: Processes, Players, and Human Impact - Nature
    There are two distinct steps of nitrification that are carried out by distinct types of microorganisms. The first step is the oxidation of ammonia to nitrite, ...
  46. [46]
    Nitrification in acidic and alkaline environments - PMC
    During aerobic chemolithoautotrophic nitrification, microorganisms convert ammonia (NH3) to nitrite (NO2−) and then nitrate (NO3−) [1]. The first step is ...
  47. [47]
    Nitrification - Arp - Major Reference Works - Wiley Online Library
    Sep 15, 2009 · Ammonia-oxidizing bacteria first oxidize ammonia to nitrite, followed by the oxidation of nitrite to nitrate by the nitrite-oxidizing bacteria.
  48. [48]
    Nitrification - an overview | ScienceDirect Topics
    The first step of nitrification is carried out by the ammonia-oxidizing bacteria (AOB) that oxidize ammonia to nitrite. Due to the slow growth of these bacteria ...
  49. [49]
    [PDF] Nitrification - Princeton University
    Nitrification is the biological conversion of ammonium to nitrate, essential in the nitrogen cycle of soils, waters, and wastewater.<|separator|>
  50. [50]
    Facilitating Nitrification Inhibition through Green, Mechanochemical ...
    Sep 13, 2021 · In soils and water, nitrapyrin inhibits the activity of ammonia monooxygenase, a microbial enzyme that catalyzes the first step of nitrification ...
  51. [51]
    Effect of nitrification inhibitors on the growth and activity of ...
    DCD and nitrapyrin inhibited nitrification at similar concentrations to bacterial ammonia oxidisers. Although DCD completely inhibited nitrification, some ...
  52. [52]
    The Role Of Nitrogen Fertilizer In Soil pH Levels - No-Till Farmer
    Feb 6, 2013 · Hydrogen (H+) is released in the process of nitrification, and free hydrogen ions increase acidity. The higher the percentage of ammonium ...
  53. [53]
    N fertilization release protons during nitrification, could someone ...
    Apr 5, 2019 · Two hydrogen ions (i.e. H+) (referred to as protons) are released following each nitrification (ammoniacal-N to nitrate-N) process.<|control11|><|separator|>
  54. [54]
    Complete nitrification by Nitrospira bacteria - Nature
    Nov 26, 2015 · Discovery of comammox. The obtained bacterial genomes were screened for the key functional genes of autotrophic nitrification. As expected ...
  55. [55]
    Metagenomic Evidence for the Presence of Comammox Nitrospira ...
    Dec 30, 2015 · The recently published discovery of a comammox Nitrospira organism (5, 6) in combination with our finding strongly suggests that microbial ...
  56. [56]
    Global patterns and drivers of coupling between anammox ... - Nature
    Jan 11, 2025 · Denitrification and anammox collectively drive nitrogen loss from aquatic ecosystems, yet their global patterns and interactions remain unclear.
  57. [57]
    Specific Denitrifying and Dissimilatory Nitrate Reduction to ... - NIH
    Jan 20, 2022 · Both denitrifying (DN) and dissimilatory nitrate reduction to ammonium (DNRA) bacteria can potentially consume excess nitrite in an anammox system.
  58. [58]
    Estimating global terrestrial denitrification from measured N2O:(N2O ...
    We estimate that terrestrial denitrification has doubled since pre-industrial times and is now in the range of 115–202 Tg-N year−1, removing 56% of the newly ...
  59. [59]
    Understanding Global Warming Potentials | US EPA
    Jan 16, 2025 · Nitrous Oxide (N2O) has a GWP 273 times that of CO2 for a 100-year timescale. N2O emitted today remains in the atmosphere for more than 100 ...
  60. [60]
    [PDF] IPCC Global Warming Potential Values - GHG Protocol
    Aug 7, 2024 · IPCC Global Warming Potential (GWP) values relative to CO2. Common ... N2O. 298. 265. 273. Nitrogen trifluoride. NF3. 17,200. 16,100. 17,400.
  61. [61]
    [PDF] Dissimilatory nitrate reduction to ammonium (DNRA) is marginal ...
    40. (DNRA). Denitrification and DNRA are competitive pathways and hence it is critical to evaluate their functional biogeochemical role.
  62. [62]
    Dissimilatory Nitrate Reduction to Ammonium (DNRA) and ...
    Jan 4, 2021 · Previous and current studies have shown that DNRA is favored under higher ambient carbon-to-nitrogen (C/N) ratios, whereas denitrification is ...
  63. [63]
    Competitive Roles of DNRA and Denitrification on Organic Nitrogen ...
    Apr 23, 2025 · Results from kinetic modeling further illustrate the competitive dynamics between DNRA and denitrification.
  64. [64]
    The competition between heterotrophic denitrification and DNRA ...
    Oxygen is an important factor affecting the competition between DEN and DNRA. Both DEN and DNRA prefer oxygen-limited environments. However, most denitrifying ...
  65. [65]
    Anaerobic ammonium oxidation (anammox) in the marine environment
    Anammox, anaerobic ammonium oxidation with nitrite, is now recognized as an important process in the marine nitrogen cycle.
  66. [66]
    Biogeography of anaerobic ammonia-oxidizing (anammox) bacteria
    After being discovered in a wastewater treatment plant (WWTP), anammox bacteria were subsequently characterized in natural environments, including marine, ...
  67. [67]
    [PDF] Significance of anaerobic ammonium oxidation in the ocean
    The anammox bacterium has yet to be isolated, but has nevertheless been described [1–3] and detected in the water column of two marine environments, both by its ...
  68. [68]
    A twilight for the complete nitrogen removal via synergistic partial ...
    Jun 14, 2021 · This research opens the opportunity of complete nitrogen removal via synergistic partial-denitrification, anammox, and DNRA (SPDAD) process.
  69. [69]
    Nitrate formation in anammox process: Mechanisms and operating ...
    As described in section 2.2, anammox, denitrification and DNRA are involved in nitrate formation in anammox process, corresponding to functional ...
  70. [70]
    Diversity and Activity of Free-Living Nitrogen-Fixing Bacteria ... - NIH
    Most (∼80%) of biological nitrogen fixation (BNF) is carried out by diazotrophs in symbiosis with legumes (33). However, under specific conditions bacteria ...
  71. [71]
    Biological Nitrogen Fixation | Learn Science at Scitable - Nature
    Many heterotrophic bacteria live in the soil and fix significant levels of nitrogen without the direct interaction with other organisms. Examples of this type ...
  72. [72]
    Plants with arbuscular mycorrhizal fungi efficiently acquire Nitrogen ...
    Mar 28, 2022 · We investigated the role of plants and their plant-derived carbon in shaping the microbial community that decomposes substrates and traced ...
  73. [73]
    Litter quality, mycorrhizal association, and soil properties regulate ...
    Feb 1, 2022 · Mycorrhizal association was also found to have direct effects on microbial communities and their activities via litter quality and indirectly ...
  74. [74]
    Influence of Arbuscular Mycorrhizal Fungi on Nitrogen Dynamics ...
    Jan 13, 2025 · Arbuscular mycorrhizal fungi (AMF) can preferentially absorb the released ammonium (NH4+) over nitrate (NO3−) during litter decomposition.
  75. [75]
    Changing perspectives on terrestrial nitrogen cycling: The ...
    Jun 2, 2019 · Our review explores how ecological and evolutionary interactions and feedbacks among N inputs, soil N and organisms (both plants and soil ...
  76. [76]
    Forest ecosystem response to four years of chronic nitrate and ...
    In general, N leaching losses increased with increasing N additions. Nitrogen retention ranged from 93 to 97% of inputs in the control and N amended sites.
  77. [77]
    Nitrogen leaching in response to increased nitrogen inputs in ...
    Under ambient condition, DIN losses in the leaching solution in the old-growth forest were 38.7–35.5 kg N ha−1 yr−1, significantly higher (P < 0.001, two-way ...
  78. [78]
    Implementation of mycorrhizal mechanisms into soil carbon model ...
    Mar 14, 2022 · Our results suggest that mycorrhizal impacts on litter decomposition are underpinned by distinct decomposition pathways in AM- and EM-dominated ecosystems.<|separator|>
  79. [79]
    CENTURY: Modeling Ecosystem Responses to Climate Change ...
    Apr 26, 2024 · The CENTURY model, Version 4, is a general model of plant-soil nutrient cycling that is being used to simulate carbon and nutrient dynamics ...Missing: CN | Show results with:CN
  80. [80]
    Modelling C and N dynamics in forest soils with a modified version ...
    The CENTURY model of soil organic matter turn-over developed by Parton and co-workers has been used successfully for grasslands to predict dynamics of C, N and ...
  81. [81]
    Tipping the plant-microbe competition for nitrogen in agricultural soils
    Oct 18, 2024 · In this short review, we will detail several examples of plant-microbe and microbe-microbe competition for nitrogen, namely amino acid uptake, ...
  82. [82]
    Denitrification across landscapes and waterscapes: A synthesis
    Freshwater systems (groundwater, lakes, rivers) account for about 20% and estuaries 1% of total global denitrification. Denitrification of land-based N sources ...Missing: percentage | Show results with:percentage
  83. [83]
    Global nitrogen budget revisited - Science
    Jun 12, 2025 · Nitrogen produced from inland waters accounts for ~10% of the global annual fixed nitrogen by all natural processes—including lightning and ...<|separator|>
  84. [84]
    Global importance of nitrogen fixation across inland and coastal ...
    Jun 12, 2025 · Inland and coastal aquatic systems fix 40 (30 to 54) teragrams of nitrogen per year, equivalent to 15% of the nitrogen fixed on land and in the open ocean.Missing: pre- | Show results with:pre-
  85. [85]
    Longer duration of seasonal stratification contributes to widespread ...
    Dec 6, 2022 · Longer duration of seasonal stratification contributes to widespread increases in lake hypoxia and anoxia · DATA AVAILABILITY STATEMENT.
  86. [86]
    Climate-driven deoxygenation of northern lakes - Nature
    Jul 1, 2024 · Seasonal hypoxia tends to develop more rapidly in stratified lakes with high nutrient concentrations20 and a large sediment surface area ...
  87. [87]
    The marine nitrogen cycle: recent discoveries, uncertainties and the ...
    Because the global estimate of 400 Tg N yr−1 removal per year exists, about 200–300 Tg N yr−1 remain for the open ocean (figure 2). Recent estimates ...
  88. [88]
    UPTAKE OF NEW AND REGENERATED FORMS OF NITROGEN IN ...
    Rates of nitrogen fixation are found to be as high or higher than nitrate uptake, in some cases suggesting an important role for nitrogen-fixing phytoplankton.
  89. [89]
    Advances in Understanding the Marine Nitrogen Cycle in the ...
    May 1, 2024 · At a global scale, higher N2 fixation rates in the Pacific (95–100 Tg N yr–1) than in the Atlantic (20–34 Tg N yr–1) and the Indian Oceans (22– ...
  90. [90]
    Amount of nitrogen fixed annually - Biosphere - BNID 105551
    Current estimates of global N2 fixation are ~240 TgN/y with a marine contribution of 100–190 TgN/y. Of this, a single nonheterocystous cyanobacteria genus ...<|separator|>
  91. [91]
    The marine nitrogen cycle: new developments and global change
    Feb 7, 2022 · Ocean warming and acidification are accompanied by other anthropogenic stressors that can be equally disruptive to the marine nitrogen cycle.
  92. [92]
    Small sinking particles control anammox rates in the Peruvian ...
    May 28, 2021 · Up to 40% of the ocean's fixed nitrogen is lost in oxygen minimum zones (OMZs) by anammox, but despite the importance of this process, nitrogen ...
  93. [93]
    Massive nitrogen loss from the Benguela upwelling system ... - PNAS
    Here, we show that instead, the anammox process (the anaerobic oxidation of ammonium by nitrite to yield N2) is mainly responsible for nitrogen loss in the OMZ ...<|separator|>
  94. [94]
    Changing perspectives in marine nitrogen fixation - Science
    May 15, 2020 · Marine N2 fixation, along with other components of the marine nitrogen cycle, will be affected by these stressors. Upper-ocean warming may ...
  95. [95]
    Global declines in oceanic nitrification rates as a consequence of ...
    Dec 20, 2010 · Our results suggest that ocean acidification could reduce nitrification rates by 3–44% within the next few decades, affecting oceanic nitrous oxide production.
  96. [96]
    Response of N2O production rate to ocean acidification in the ...
    May 11, 2020 · Collectively these results suggest that, if seawater pH continues to decline at the same rate, ocean acidification could increase the marine N2O ...
  97. [97]
    origin and evolution of Earth's nitrogen | National Science Review
    However, it is estimated that the early atmosphere of Earth may have contained ∼1.4 times the present-day atmospheric nitrogen (PAN), with ∼0.4 PAN being ...
  98. [98]
    The global nitrogen cycle in the twenty-first century - Journals
    Jul 5, 2013 · [16] provides an estimate of annual pre-industrial BNF in terrestrial ecosystems of 58 Tg N, within a range of 40–100 Tg, and a discussion of ...
  99. [99]
    Review Research progress of urban nitrogen cycle and metabolism
    The average annual nitrogen fixation rate has reached 150 Tg N/yr, which is nearly 2.5 times the Earth's safe boundary value (62 Tg N/yr) and is still ...
  100. [100]
    Global net climate effects of anthropogenic reactive nitrogen - Nature
    Jul 24, 2024 · Anthropogenic NOx emissions during 2016–2020 reached 46.5 ± 2.7 TgN yr−1, most of which were due to fossil fuel combustion, as derived from CEDS ...Missing: pool | Show results with:pool
  101. [101]
    Convergent evidence for widespread rock nitrogen sources in ...
    Apr 6, 2018 · Here we demonstrate that bedrock is a nitrogen source that rivals atmospheric nitrogen inputs across major sectors of the global terrestrial environment.
  102. [102]
    Human-caused increases in reactive nitrogen burial in sediment of ...
    By synthesizing records of Nr burial in sediments from 303 lakes worldwide, here we show that 9.6 ± 1.1 Tg N year−1 (Tg = 1012 g) accumulated in inland water ...
  103. [103]
    Human-caused increases in reactive nitrogen burial in sediment of ...
    Aug 12, 2021 · Considering the contribution from natural sources, we estimate a 6.6 Tg N year−1 burial in global lakes derived from anthropogenic sources, ...
  104. [104]
    Global soil metagenomics reveals distribution and predominance of ...
    May 24, 2024 · We suspect that nifD/K, rather than nifH, serve as a reliable marker gene for nitrogen-fixing populations (especially in shotgun metagenomic ...Missing: 2020s | Show results with:2020s
  105. [105]
    Detecting Nitrous Oxide Reductase (nosZ) Genes in Soil ...
    The present study revealed that atypical nosZ genes outnumbered their typical counterparts, highlighting their potential role in N2O consumption in soils and ...
  106. [106]
    Genome-resolved metagenomics reveals abundant nitrate reducers ...
    Mar 2, 2023 · From 56 metagenomes spanning all three major ODZs, we use genome-resolved metagenomics to reveal the predominance of partial denitrifiers, particularly single- ...Gene Searching Within... · Results · Denitrifying Phyla And...
  107. [107]
    Novel metagenome-assembled genomes involved in the nitrogen ...
    Jun 18, 2021 · Here we recovered 39 high- and medium-quality metagenome-assembled genomes (MAGs) from the Eastern Tropical South Pacific OMZ.
  108. [108]
    (PDF) Nitrogen Cycle in Agriculture: Biotic and Abiotic Factors ...
    Nov 4, 2021 · Factors such as soil pH, oxygen availability, temperature, and microbial composition regulate nitrogen transformations and losses. ...
  109. [109]
    Consequence of altered nitrogen cycles in the coupled human and ...
    Information on the processes influencing N2O emissions from soils is sparse, particularly that on the roles of temperature, moisture, redox potential, pH, and ...
  110. [110]
    [PDF] Human Alteration of the Global Nitrogen Cycle: Causes and ...
    Industrial fixation of nitrogen for use as fertilizer currently totals approximately 80 Tg per year and repre- sents by far the largest human contribution of ...
  111. [111]
    Regulation of Redfield ratios in the deep ocean - AGU Publications
    Jan 23, 2015 · Nitrogen and phosphorus oceanic cycles are heavily affected by anthropogenic activities, mainly through an increase in N and P supply by ...
  112. [112]
    Unraveling the genetic potential of nitrous oxide reduction in ...
    Aug 13, 2024 · This study provides critical insights into the genetic diversity of nitrous oxide reductase (nosZ) genes and the microorganisms harboring them ...
  113. [113]
    Interactive effects of soil temperature and moisture on soil N ...
    Apr 1, 2014 · The results showed that the mineral N production and net N mineralization rates were positively correlated with temperature; the strongest ...
  114. [114]
    Soil net nitrogen mineralisation across global grasslands - Nature
    Oct 31, 2019 · Further, potential soil net Nmin was positively influenced by temperature seasonality and was negatively affected by temperature of the wettest ...
  115. [115]
    Precipitation affects soil nitrogen fixation by regulating active ...
    Apr 1, 2024 · Our study suggested that appropriate precipitation reduction can enhance soil N fixation through promoting the abundance of the soil active diazotrophs and ...
  116. [116]
    Drought Stress Modifies the Source‐Sink Dynamics of Nitrogen ...
    May 26, 2025 · Under severe water deficit conditions, almost complete inhibition of nitrogen fixation was observed, with an 88.8% decline (Table 1). These ...
  117. [117]
    Nitrogen and Phosphorus Limitation over Long-Term Ecosystem ...
    Aug 3, 2012 · If NPP switches from N to P limitation over long time periods, the transition time depends most strongly on the P weathering rate. At all ...
  118. [118]
    Evolving Nitrogen-Fixing Legume Symbionts - PubMed
    The ability to fix nitrogen with legumes is a remarkable example of a complex trait spread by horizontal transfer of a few key symbiotic genes.
  119. [119]
    Humboldt Review: Are legumes different? Origins and ...
    This review examines the evolutionary journey that has led to the diversification of legumes, in particular its nitrogen-fixing symbiosis.
  120. [120]
  121. [121]
    A model for large glacial–interglacial climate-induced changes in ...
    The model estimates a southern South America dust source activity two to three times higher for glacial periods than for the Holocene and a glacial temperature ...
  122. [122]
    The Haber Process - Chemistry LibreTexts
    Jan 29, 2023 · General Conditions of the Process. The catalyst: The catalyst is actually slightly more complicated than pure iron.
  123. [123]
    Haber-Bosch Process - Stanford University
    Dec 14, 2022 · The reaction runs at temperatures in the range of 400°C-500 °C, pressure in the range of 150-300 bar, and with the presence of an iron-based ...
  124. [124]
    Reaction Mechanisms, Kinetics, and Improved Catalysts for ...
    Apr 6, 2022 · The Haber–Bosch (HB) process is the primary chemical synthesis technique for industrial production of ammonia (NH3) for manufacturing ...
  125. [125]
    The Haber Bosch Catalyst from Solid state Chemistry to ...
    Jun 19, 2025 · It self-assembles from an oxide precursor during reductive activation into an intricate and hierarchical porous entity with unique properties, ...Introduction · Results · Discussion · Experimental Section
  126. [126]
    The industrialization of the Haber-Bosch process - ACS Publications
    Aug 14, 2023 · Since the start of this decade, they have produced an average of over 10,000 kg of corn per hectare. A 2008 study in Nature Geoscience ...
  127. [127]
    Executive Summary – Ammonia Technology Roadmap – Analysis
    Ammonia production currently relies heavily on fossil fuels. Global ammonia production today accounts for around 2% (8.6 EJ) of total final energy consumption.Missing: H2 2020s
  128. [128]
    Synthetic Nitrogen Fertilizer in the U.S. - farmdoc daily
    Feb 17, 2021 · The discovery of the Haber-Bosch process allowed for the widespread fertilization of crops, and together with other agricultural technology ...
  129. [129]
    The Haber-Bosch Process - Institute for Mindful Agriculture
    Mar 24, 2016 · The introduction of nitrogenous fertilizers and their increasing application had a dramatic impact on grain yields. Together with new high ...
  130. [130]
    Technology status: ammonia production from electrolysis-based ...
    Jan 31, 2023 · By the end of this year, electrolysis-based ammonia production is estimated to account for 185,000 tonnes per year.Missing: H2 adoption rate 2020s
  131. [131]
    Operando probing of the surface chemistry during the Haber–Bosch ...
    Jan 10, 2024 · The active component of the catalyst enabling the conversion was variously considered to be the oxide, nitride, metallic phase or surface ...
  132. [132]
    Types and Uses of Nitrogen Fertilizers for Crop Production
    Fertilizers common to crop production in Indiana usually contain nitrogen in one or more of the following forms: nitrate, ammonia, ammonium or urea. Each form ...
  133. [133]
    Nitrate vs Ammonium vs Urea: Nitrogen Fertilizers Forms - Yarafert
    The three primary forms of nitrogen fertilizer used globally are nitrate (NO 3 – ), ammonium (NH 4 + ), and urea (CO(NH 2 ) 2 ).
  134. [134]
    Nitrogen and the future of agriculture: 20 years on | Ambio
    Mar 14, 2021 · Current annual global fertilizer N use on cropland is about 100 Tg (excluding grasslands and forage crops). N use efficiency is a broader ...
  135. [135]
    World Cereal Nitrogen Use Efficiency Trends: Review and Current ...
    Jun 20, 2019 · Cereal world NUE was estimated and reported by Raun and Johnson (1999) at 33%. This implies that 67% of all the applied N is unaccounted for and ...
  136. [136]
    [PDF] Ammonia volatilization from synthetic fertilizers and its ... - Yiqi Luo
    We found that the global average percentage of N loss as NH3 was 17.6%, which is comparable to the range of 10–14% reported by the IPCC (De Klein et al., 2006) ...
  137. [137]
    The Green Revolution: Norman Borlaug and the Race to Fight ... - PBS
    Apr 22, 2025 · Borlaug's wildly successful efforts to increase crop yields came to be known as the “Green Revolution” and earned him the Nobel Peace Prize ...
  138. [138]
    Benefits of Increasing Information Accuracy in Variable Rate ...
    Variable rate technologies allow the right quantities of fertilizer to be applied at the right place. This helps to both maintain yields and avoid nitrogen ...Methodological And... · 1. Introduction · 5. Results
  139. [139]
    UW innovations are helping farmers produce crops with less ...
    Jun 11, 2025 · UW researchers are engineering beneficial bacteria that produce a process called nitrogen fixation in crops to improve their resilience.Missing: microbiome 2020s
  140. [140]
    Engineering nitrogen-fixing microbiomes with waste-derived carbon ...
    Sep 23, 2025 · This study employed microbiome engineering to develop a self-assembled nitrogen-fixing microbial community utilizing carbon compounds from ...2.2. Microbiome Engineering... · 2.3. Microbiome Engineering... · 3. ResultsMissing: 2020s | Show results with:2020s
  141. [141]
  142. [142]
    Global nitrous oxide budget (1980–2020) - ESSD Copernicus
    Jun 11, 2024 · According to the BU approaches, the total amount of global natural N2O emissions fluctuated between 11.7 and 12.1 Tg yr−1 during 1980–2020.Missing: sewage | Show results with:sewage
  143. [143]
    Nitrogen cycling during wastewater treatment - ScienceDirect.com
    On a global scale humans excrete about 23 Tg of reactive nitrogen per year (Smil, 1999) and more than half of this waste is released into the environment ...
  144. [144]
    Nitrous oxide emissions from wastewater treatment processes - PMC
    Global N2O emissions from wastewater treatment are expected to increase by approximately 13 per cent between 2005 and 2020. N2O is mainly released during ...
  145. [145]
    Global nitrogen and phosphate in urban wastewater for the period ...
    Sep 11, 2009 · This paper presents estimates for global N and P emissions from sewage for the period 1970–2050 for the four Millennium Ecosystem Assessment ...<|separator|>
  146. [146]
    4.2.3.3 Nitrogen oxides (NO x ) - IPCC
    The NOx emissions from fossil fuel use used in model studies here for year 2000 are considerably higher, namely 33 TgN/yr. The large American and European ...<|separator|>
  147. [147]
    How many people does synthetic fertilizer feed? - Our World in Data
    Nov 7, 2017 · This amounts to 44 percent of the global population in 2000 being fed by nitrogen fertilizers, rising to 48 percent in 2008. Here we have ...
  148. [148]
    Nitrogen fertilizer: Retrospect and prospect - PMC - PubMed Central
    During the 1960s, the 4.3% rise per yr of GDP per person plus the population increase raised world GDP faster than 6% per yr. Then the change in GDP per person ...
  149. [149]
    [PDF] Benefits of nitrogen for food, fibre and industrial production
    Farmers have an economic incentive to apply only the economically optimal rate of fertilizer N, but there is no strong incentive to increase N use efficiency ...
  150. [150]
    Gulf of Mexico 'dead zone' larger than average, scientists find - NOAA
    Aug 1, 2024 · The resulting low oxygen levels (hypoxia) cause animals, like fish and shrimp, to leave the area. Exposure to hypoxic waters has been found to ...
  151. [151]
    Gulf Dead Zone | The Nature Conservancy
    Feb 18, 2025 · What Causes the Dead Zone? Heavy rains and melting snows wash massive amounts of nutrients—particularly nitrogen and phosphorus—from lawns, ...
  152. [152]
    Nitrogen‐induced terrestrial eutrophication: cascading effects and ...
    Jul 31, 2017 · We found 21 N critical loads were triggered by N deposition (ranging from 2 to 39 kg N·ha−1·yr−1), which cascaded to distinct beneficiary types ...Abstract · Methods · Results · Discussion
  153. [153]
    A global analysis of soil acidification caused by nitrogen addition
    We found that N addition significantly reduced soil pH by 0.26 on average globally. However, the responses of soil pH varied with ecosystem types, N addition ...Missing: drop | Show results with:drop
  154. [154]
    Soil Acidification Can Be Improved under Different Long-Term ...
    Jun 24, 2024 · Especially in the treatment of applying nitrogen fertilizer, the acidification rate reaches 0.121~0.225 units per year, which is much higher ...
  155. [155]
    Estimated Nitrate Concentrations in Groundwater Used for Drinking
    Jun 5, 2025 · While nitrate does occur naturally in groundwater, concentrations greater than 3 mg/l generally indicate contamination (Madison and Brunett, ...Missing: leaching | Show results with:leaching
  156. [156]
    Impact of nitrogen deposition at the species level - PNAS
    Here, we use a large dataset of European grasslands along a gradient of nitrogen (N) deposition to show statistically significant declines in the abundance of ...Missing: native | Show results with:native
  157. [157]
    Ecological thresholds under atmospheric nitrogen deposition for ...
    Alliance-level thresholds (change points) for species decreasing in abundance along the gradient ranged from 1.8−14.3 kg N ha−1 yr−1 and tended to be lower in ...
  158. [158]
    Vegetation community change points suggest that critical loads of ...
    Community-level change points indicated a decrease in species abundance along a nitrogen deposition gradient ranging from 3.9 to 15.3 kg N ha−1 yr−1, which were ...
  159. [159]
    Overview of Greenhouse Gases | US EPA
    Jan 16, 2025 · Greenhouse gases are gases that trap heat in the atmosphere. Main types include carbon dioxide, methane, nitrous oxide, and fluorinated gases.
  160. [160]
    Nitrous oxide emissions grew 40 percent from 1980 to 2020 ...
    Jun 12, 2024 · Emissions of nitrous oxide, the third most important human-made greenhouse gas, rose 40 percent from 1980 to 2020, according to a new report by the Global ...
  161. [161]
    Measuring and modeling the lifetime of nitrous oxide including its ...
    The result is 116 ± 9 years. The observed monthly‐to‐biennial variations in lifetime and tropical abundance are well matched by four independent chemistry‐ ...
  162. [162]
    Ground-level Ozone Basics | US EPA
    Mar 11, 2025 · Ozone at ground level is a harmful air pollutant, because of its effects on people and the environment, and it is the main ingredient in “smog."
  163. [163]
    Attributing ozone to NOx emissions: Implications for climate ...
    Ozone is an important greenhouse gas. Hence NOx related ozone changes have an impact on climate that is more complex than climate effects of CO2 increases.
  164. [164]
    Global net climate effects of anthropogenic reactive nitrogen - PMC
    Jul 24, 2024 · This increase can be mainly attributed to emissions associated with anthropogenic fossil fuel combustion and fertilizer application.
  165. [165]
    Changing patterns of global nitrogen deposition driven by socio ...
    Jan 2, 2025 · This database show that global nitrogen deposition to land is 92.7 Tg N in 2020. ... Reactive N emissions from cropland and their mitigation in ...
  166. [166]
    Exploring global changes in agricultural ammonia emissions ... - PNAS
    Mar 28, 2022 · Global NOy deposition on land has increased by 40% from 1980 (27 Tg N) to 2016 (38 Tg N), after then it has declined by 16% to 2018 (32 Tg N).
  167. [167]
    Nitrogen deposition and climate change effects on tree species ...
    Dec 10, 2018 · Elevated N deposition can have both nutrient enrichment and acidifying impacts, and the resulting effects on trees within forest ecosystems can ...
  168. [168]
    The impact of nitrogen deposition on carbon sequestration by ...
    The increase in N deposition on forests may increase C sequestration by increased growth and increased accumulation of soil organic matter (SOM) through ...
  169. [169]
    Changing patterns of global nitrogen deposition driven by socio ...
    Jan 2, 2025 · This database show that global nitrogen deposition to land is 92.7 Tg N in 2020. Total nitrogen deposition increases initially, stabilizing after peaking in ...
  170. [170]
    Chemical Contaminant Rules | US EPA
    There is an acute health risk from nitrate and nitrite. The regulations reduce the risk of Methemoglobinemia or "blue baby syndrome." Blue Baby Syndrome is ...
  171. [171]
    Executive Summary - Nitrate and Nitrite in Drinking Water - NCBI
    Infection is the major contributor to methemoglobinemia from nitrate exposure; the incremental contribution of drinking water is negligible. There are very few ...
  172. [172]
    Risk assessment of N‐nitrosamines in food - - 2023 - EFSA Journal
    Mar 28, 2023 · ... classified by the International Agency for Research on Cancer (IARC) as probably carcinogenic to humans (Group 2A). The EFSA Panel on Food ...
  173. [173]
    [PDF] IARC Monographs on the Evaluation of Carcinogenic Risks to ...
    N-nitrosamines can also form in the stomach unless inhibited by vitamin C or other antioxidants.
  174. [174]
    The Respiratory Risks of Ambient/Outdoor Air Pollution - PMC
    The World Health Organization estimates that ambient pollution is responsible for 4.2 million premature deaths annually. Several studies in the developed ...
  175. [175]
    Ambient (outdoor) air pollution - World Health Organization (WHO)
    Oct 24, 2024 · Ambient (outdoor) air pollution is estimated to have caused 4.2 million premature deaths worldwide in 2019. Some 89% of those premature deaths ...
  176. [176]
    Reducing the Health Impacts of the Nitrogen Problem - NCBI - NIH
    Sep 30, 2021 · Too much nitrate consumption can pose a health risk. Users of public drinking water supplies and private wells in areas surrounded by farmland ...