Next Article in Journal
Estimation of Nitrogen Concentration in Walnut Canopies in Southern Xinjiang Based on UAV Multispectral Images
Next Article in Special Issue
Design and Implementation of Evaluation Method for Spraying Coverage Region of Plant Protection UAV
Previous Article in Journal
Evaluation of the Intra- and Interspecific Development of Different Accessions of Silphium perfoliatum L. and Silphium integrifolium Michx.
Previous Article in Special Issue
Effect Application of Apple Pomace on Yield of Spring Wheat in Potting Experiment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Deciphering the Interactions in the Root–Soil Nexus Caused by Urease and Nitrification Inhibitors: A Review

1
School of BioSciences, University of Melbourne, Parkville, VIC 3010, Australia
2
School of Chemistry, Bio21 Institute, The University of Melbourne, 30 Flemington Road, Parkville, VIC 3010, Australia
3
Research School of Biology, The Australian National University, Acton, ACT 2601, Australia
*
Author to whom correspondence should be addressed.
Agronomy 2023, 13(6), 1603; https://doi.org/10.3390/agronomy13061603
Submission received: 20 May 2023 / Revised: 9 June 2023 / Accepted: 11 June 2023 / Published: 13 June 2023

Abstract

:
Optimizing nitrogen (N) availability to plants is crucial for achieving maximum crop yield and quality. However, ensuring the appropriate supply of N to crops is challenging due to the various pathways through which N can be lost, such as ammonia (NH3) volatilization, nitrous oxide emissions, denitrification, nitrate (NO3) leaching, and runoff. Additionally, N can become immobilized by soil minerals when ammonium (NH4+) gets trapped in the interlayers of clay minerals. Although synchronizing N availability with plant uptake could potentially reduce N loss, this approach is hindered by the fact that N loss from crop fields is typically influenced by a combination of management practices (which can be controlled) and weather dynamics, particularly precipitation, temperature fluctuations, and wind (which are beyond our control). In recent years, the use of urease and nitrification inhibitors has emerged as a strategy to temporarily delay the microbiological transformations of N-based fertilizers, thereby synchronizing N availability with plant uptake and mitigating N loss. Urease inhibitors slow down the hydrolysis of urea to NH4+ and reduce nitrogen loss through NH3 volatilization. Nitrification inhibitors temporarily inhibit soil bacteria (Nitrosomonas spp.) that convert NH4+ to nitrite (NO2), thereby slowing down the first and rate-determining step of the nitrification process and reducing nitrogen loss as NO3 or through denitrification. This review aims to provide a comprehensive understanding of urease and nitrification inhibitor technologies and their profound implications for plants and root nitrogen uptake. It underscores the critical need to develop design principles for inhibitors with enhanced efficiency, highlighting their potential to revolutionize agricultural practices. Furthermore, this review offers valuable insights into future directions for inhibitor usage and emphasizes the essential traits that superior inhibitors should possess, thereby paving the way for innovative advancements in optimizing nitrogen management and ensuring sustainable crop production.

1. Introduction

Nitrogen (N) is essential to the survival of all living organisms [1]. The global N cycle involves the flow of atmospheric nitrogen (N2) into terrestrial and marine ecosystems through N fixation, a reductive process that produces ammonia (NH3) or ammonium (NH4+), which is converted by microbiological processes into higher oxidized forms of nitrogen (nitrification) and finally returns to the atmosphere as reduced, gaseous forms of nitrogen (denitrification) [2]. In contrast to the unreactive atmospheric N2, the various reduced or oxidized forms of N are often described as “reactive N” (Nr) [3]. Despite the abundance of N2 in the environment, Nr is a major limitation to global net primary productivity and food production [4]. Increasing the availability of Nr is therefore essential for plant growth and food production for the continuously growing world population.
The invention of the Haber–Bosch process in the early 20th century, which enabled the catalytic hydrogenation of N2 under elevated pressure and temperature to produce biologically available NH3, triggered the production of N fertilizers on a large industrial scale [5]. The Haber–Bosch process is considered as the most important industrial development of the 20th century, which was recognized by two Nobel Prizes in chemistry awarded to Fritz Haber in 1918 and Carl Bosch in 1931 [6]. Industrially fixed N fertilizers revolutionized agricultural productivity worldwide, mainly through an application on staple grain crops, wheat, rice, and maize [7]. Nowadays, about 80% of the total Haber–Bosch produced N goes to plant production [8,9].
With the expected world population to hit the 10 billion mark by 2050, the reliance on industrial N fertilization will significantly increase further. While the Haber–Bosch process itself is optimized and has nearly reached thermodynamic process efficiency [10], unfortunately, the application of Haber–Bosch N in agriculture is highly inefficient. Urea is the dominating N fertilizer worldwide [11], which has a N content of 46%. Unfortunately, a high proportion (50–70% of applied Nr) is not taken up by the plant roots and “lost” to the environment [12,13], leading to the accumulation of nitrate (NO3) in terrestrial and aquatic ecosystems, resulting in the contamination of drinking water and eutrophication of water bodies, which has negative impacts for many ecosystems and significant consequences for the earth’s climate and environmental health [8,14,15,16]. In addition, the emission of gaseous Nr species, such as NH3, which is a precursor to potentially harmful microscopic particulate matter (PM2.5) [17,18,19], and the greenhouse gas nitrous oxide (N2O) into the troposphere contributes to air pollution, global warming, and the depletion of stratospheric ozone [19]. Hence, the effectiveness of N fertilizers must be significantly increased to limit environmental degradation resulting from agricultural processes.
One of the current approaches to reduce unwanted N losses to the environment and increase plant uptake is by slowing down the rate of urea hydrolysis in soils using urease inhibitors and by slowing down the subsequent microbiological nitrification process using nitrification inhibitors. The objective of this review is to provide an overview of the current knowledge of the spatial and temporal interplay of urease and nitrification inhibitors with plants and N uptake by roots, with particular emphasis on the need to develop design principles for inhibitors with improved efficiency.

2. Nitrogen as an Important Plant Nutrient

Nitrogen (N) is the most significant macro-nutrient element, which is included in a considerable number of nutritional and energy substrates [20]. N is crucial for the growth and development of plants, as well as for metabolic activities, such as photosynthesis, energy production, biomass production, and yield [21,22]. The enzymatic activity, chlorophyll content, photosynthesis, respiration rate, and yield of crop plants are all significantly reduced by N deficiency. Therefore, sufficient N fertilizer along with inhibitors has been added to farmland soil all over the world for decades to ensure crop production. To optimize the performance of inhibitors in agricultural systems to improve N availability with plant demand, it is necessary to understand the mechanisms by which plants influence microbial N transformations to improve nutrition, identify the N forms that plants can take up, and how N compounds are influenced by the availability of various forms of N in 89 soils and by plant properties (Figure 1). Soluble N in the form of fertilizer in which nitrogen is easily accessible for plant uptake, in soil, varies greatly over space and time, and plants are capable of exploiting any ephemeral microscale patches of N [23]. Generally, plants take up N in their inorganic forms as NO3 and NH4+ [24] but are also capable to use N bound in organic matter [25].
Plant properties such as genotype act mainly via characteristics linked to transporters capable of transporting different forms of N [26] and those are susceptible to NH3 toxicity [27]. Additionally, plant N status and plant growth stage can alter rates of both NO3 and NH4+ uptake differentially [28]. Microorganisms can influence the uptake of N in plants, for instance, it has been found that competing neighboring plants [29] and competing soil microorganisms [30] may reduce the availability of certain N forms.
The capacity of plants to absorb certain forms of N can be enhanced by N-fixing symbiosis among α- and β-proteobacteria and legumes [31,32]. This symbiotic relationship between soil bacteria and legume roots is mediated by endosymbiotic interaction, where plants generate nodules (a new differentiated special organ) that help to fix atmospheric N through the action of the enzyme nitrogenase [33]. A review by Mahmud et al. (2020) [34] collates several studies on the mechanism of N fixation by legumes and suggests potential approaches to introduce these mechanisms into economically important crops, such as rice, maize, and wheat.

2.1. Acquisition and Assimilation of N by Plants

In general, plants take up N in its inorganic forms such as NO3 and NH4+ [24]. The NH4+ uptake is either active or passive, depending on its concentration in soils [35]. The active uptake could be achieved by direct use of metabolic energy to carry a solute across a membrane toward a region of higher electrochemical potential. The passive uptake could be completed by solute flux across a membrane along the electrochemical potential gradient, in response to a favorable electrical gradient [36]. However, NH4+ is bound to negatively charged soil particles through electrostatic interactions, reducing its mobility in soils, while NO3 is more mobile in such soil conditions and therefore principally more available for plant N uptake [24]. Uptake of NO3 is an active, energy-demanding process as it is absorbed against an electrochemical gradient [37] and requires reduction to NH4+ for assimilation within the plant, which makes it less energetically favorable than NH4+ uptake.
Several inorganic and organic N transporters have been identified in roots with different substrate affinities and specificities. Detailed mechanisms of N root uptake with inorganic N transporters for NO3 and NH4+ in several cereal crops can be found in the work of Tegeder and Masclaux-Daubresse [38]. In Arabidopsis, NH4+ acquisition by roots is accomplished by four ammonium transporters (AMTs): AMT1;1, AMT1;3, and AMT1;5 are involved in direct NH4+ uptake from the soil via the root epidermis, while AMT1;2 is expressed in cortical and endodermal cells and mediates apoplastic absorption of NH4+. Low- and high-affinity nitrate transporters mediate the NO3 uptake in soil. The low-affinity transport system allows transport in high (>0.5 mM) external nitrate concentrations, whereas the high-affinity transport system provides a capacity for nitrate absorption at low (<0.5 mM) external nitrate concentrations [39]. The majority of the nitrate transport (NRT1) family (which is also called as the nitrate transporter1/peptide transporter family, NPF) is described as low-affinity systems [40]. In Arabidopsis, six transporters are involved in root nitrate uptake as shown in [38]. Of these, chlorate-resistance protein 1 (CHL1) and NPF4.6/NRT1.2 mainly operate under high nitrate supply, whereas NRT2.1, NRT2.2, NRT2.4, and NRt2.5 function under NO3 starvation [41,42,43,44,45,46,47,48].
Plants can thrive on a wide range of dissolved organic N forms, ranging from simple, low-molecular-weight compounds, such as amino acids, nucleotides, and urea [49], to more complex polymeric material, such as proteins [50]. The uptake of organic N can differ between these simple and complex materials, and this difference can depend on various factors, such as accessibility, uptake mechanisms, and regulation and preferences [51]. These factors are elaborated here: (i) Accessibility—Simple, low-molecular-weight compounds such as amino acids are readily available in the soil solution and can be directly taken up by plant roots. They are easily absorbed through specific transporters in the root membranes [52,53]. In contrast, more complex polymeric materials such as proteins require additional steps for breakdown and assimilation. Proteins need to be enzymatically hydrolyzed into their constituent amino acids or smaller peptides before they can be taken up by plant roots [54]. (ii) Uptake mechanisms—The uptake of simple organic N compounds is often facilitated by specific transporters that recognize and transport these compounds across the root membranes [55,56]. Different transporters may have different affinities and specificities for different types of organic N compounds. In the case of polymeric materials such as proteins, specific extracellular enzymes secreted by plants or associated microbes are involved in the breakdown of proteins into smaller components. Once broken down, the resulting amino acids or peptides can be taken up by plant roots through the same transporters used for simple organic N compounds [51,57]. (iii) Regulation and preference—Plants possess regulatory mechanisms that can modulate the expression of transporters and enzymes involved in the uptake and utilization of different organic N forms. They can adjust their nutrient uptake strategies based on the availability and relative abundance of different organic N compounds in the soil. In some cases, plants may exhibit preferences for specific types of organic N compounds based on their metabolic requirements or environmental conditions [51,56,57].
Uptake of organic N is the least energy-consuming pathway for the plant, as no assimilation process is required [23]. Organic N can enter the roots through passive diffusion at high exogenous concentrations (>1 mM). Transport across the plasma membrane is generally driven by families of P-type proton ATPase (H+-ATP-ase) fueled active transporters, the genes for which appear to be constitutively expressed at high levels, irrespective of soil N supply [58]. Although these high-affinity transporters maintain cytoplasmic N concentrations at the mM level, demonstrating the efficiency of these N capture systems, as a consequence the external N concentration around the roots can be depleted to the nM level [59]. In turn, this steep concentration gradient across the plasma membrane results in the continual passive loss of organic N from the root back into the apoplast and soil, via rhizodeposition [60]. It has also been suggested that the expression of these transporters may be related to the exchange of signaling molecules between plant-growth-promoting microorganisms and roots or in the root sensing of their environment [61].
Despite evidence that roots can take up organic N, the ecological significance of this N-acquisition pathway remains controversial. While thousands of individual organic N compounds are present in the soil solution, many of these are composed of humic-rich by-products resulting from microbial breakdown and are largely unavailable to plants [62,63]. This is primarily because of the molecular complexity that these humic substances possess. They are composed of a complex mixture of aromatic and aliphatic compounds with diverse functional groups, including carboxyl, hydroxyl, and phenolic groups. This complexity makes it challenging for plants to directly utilize these compounds as a nitrogen source [64]. Another reason could be due to their size and molecular weight. Humic substances are typically large in size and have high molecular weights. Their large molecular size can limit their movement through soil pores and restrict their diffusion to plant roots. This physical limitation can make it difficult for plants to access and uptake these compounds effectively [65]. The binding and sorption capability of humic-rich by-products also reduce their availability for plant uptake, as humic substances have a strong affinity for cations and can form stable complexes with metal ions in the soil. These interactions can result in the binding and sequestration of organic nitrogen within the humic substances. Furthermore, although also smaller and less complex N-containing molecules are present in the soil, for example, glucosamine, which is likely a breakdown product of fungal cell wall chitin, they may not be widely taken up by roots or used in plant metabolism [66]. Numerous additional substances are present at very low concentrations (<1 µM), which may indicate that they have little bearing on plant N nutrition because they are so close to the net influx–efflux equilibrium point. However, it is their rate of replenishment and any potential reserves that may be present during the soil exchange phase that are crucial, not their concentration in the soil solution. For instance, it has been estimated that the soil solution amino acid pool is replenished more than 1000 times a day yet is maintained at low concentrations (ca. 20 µM) in all ecosystems [67]. This rate of cycling can be orders of magnitude faster than soil NH4+ and NO3 production rates.
Overall, while there is a good qualitative understanding of the different N forms that can be taken up by plants, the quantitative importance of organic N uptake is not yet well understood and requires further investigation to enable a better understanding of the ecology, physiology, and molecular biology of plant N nutrition. Potential approaches could include the study of soil solution dynamics of inorganic and organic N compounds, mechanistic understanding of root uptake processes, and plant N uptake under field conditions [68].

2.2. Plants Influence Microbial N Transformations in Soils

Plants respond differently to spatial and temporal variation in nutrient availability. This topic is of long-standing interest to ecology and has been addressed by recognizing the existence of an evolutionary trade-off between fast-growing exploitative and slow-growing conservative growth strategies. Trait-based approaches, such as phenotypic traits, are now commonly used to better categorize plant distribution along resource gradients in relation to these strategies. However, this focus has recently been shifted to below-ground root traits and associations with soil microbial communities and nutrient cycling processes [69]. It has been shown that soil N availability is strongly dependent on various microbial guilds that transform inert N2 into Nr species that can be taken up by plants and that plants can shape the activity and composition of the soil microbial communities [70,71]. As a plethora of microbial guilds are of importance for N availability in soil, it seems reasonable to hypothesize that plants have also evolved multiple mechanisms to acquire Nr by shaping and recruiting these N-cycling soil microbial communities. When examining the function of urease and nitrification inhibitors when administered as part of a fertilizer mix, it becomes especially critical to comprehend the interaction between N availability, plant N absorption, and microbial composition. We will elute to this below in more detail.
Plants can also limit microbial processes that would lead to loss of N through a range of mechanisms and traits. It has been shown that a relationship exists between plant growth strategies and/or plant traits, the activity of N-cycling microbes, and N retention and loss in the soil. Thus, de Vries and Bardgett [72] and Abalos et al. [73] demonstrated that, compared with slow-growing species, plants with exploitative growth strategies are associated with reduced N losses via leaching in soil and reduced microbial N2O emissions. Cantarel et al. [74] also showed that nitrification parameters were strongly driven by below-ground traits, such as specific root length (root length per unit of root biomass), root N, and plant affinity for NH4+, whereas structural equation modeling revealed that root length density (RLD) (root length per unit of soil volume) was a key trait regulating the effects of plants on N2O emissions as RLD has been shown to be important for N capture from nutrient patches, which are hotspots that trigger N2O emissions [73,74]. As an acquisitive strategy, plant traits such as high specific root length and/or high root length density [75] are commonly associated with high rates of soil resource acquisition [73,76]. Thus, it can be interpreted that plants and microbes compete for N based on the linkages between plant traits and the activity of N-cycling microbes. In fact, Moreau et al. [69] demonstrated that the abundance of bacteria capable of using NO3 as an alternative electron acceptor, rather than carbon released by the plant in the rhizosphere, was inversely related to root N uptake rate. This finding indicates that microbes performing the first step of denitrification, i.e., NO3 > NO2, can be outcompeted by plants with a high root N uptake rate. Therefore, the type of N fertilizer, as well as urease and nitrification inhibitors, is an important determining factor in N loss pathways as it directly affects the NO3/NH4+ concentration ratio in soil [77].

3. Influence of Soil Nitrogen Conversions on Nr Loss Control

NH4+-based fertilizers, which include urea ((H2N)2C=O), anhydrous NH3, ammonium sulfate ((NH4)2SO4), and ammonium nitrate (NH4NO3), are the most commonly used forms of N fertilizers applied in agriculture [78]. As mentioned above, of these, urea is the most common synthetic N fertilizer used in the world [79,80], due to its high N content (46%), relatively low cost per N unit, high water solubility, and high foliar uptake compared with other solid N fertilizers [81]. Urea is hydrolyzed in the soil to NH3 and carbon dioxide (CO2) via the intermediate unstable carbamic acid [82,83] (Scheme 1), which is catalyzed by the enzyme urease, a nickel (Ni)-dependent enzyme. Carbamic acid spontaneously hydrolyzes at physiological pH to form carbonic acid and a second molecule of ammonia. The rate of NH3 loss through volatilization is governed by fertilizer application rate, as well as soil pH, moisture, and temperature [5].
NH3 can be lost to the environment at the soil surface through volatilization or following deprotonation of NH4+ supplied through ammonium-containing fertilizers (for example, (NH4)2SO4 or NH4NO3) at high pH (Figure 2). Another loss channel for NH3 is microbiological nitrification that involves the stepwise oxidation of NH3 to NO3. The latter can leach into the groundwater or be converted into various gaseous N species (such as N2O, nitrogen oxide (NO), and N2) by reductive denitrification processes, ultimately resulting in loss of Nr from the soil–plant system [84].
N H 3 ammonia monooxygenase ( AMO ) N H 2 O H hydroxylamine oxidoreductase ( HAO ) N O 2 nitrite oxidoreductase ( NIX ) N O 3
Nitrification occurs in three steps, where, in the first step, NH3 is oxidized to hydroxylamine (NH2OH) by ammonia-oxidizing bacteria (AOB). This step is rate-limiting and catalyzed by the enzyme ammonia monooxygenase (AMO), which is encoded by the amoA gene found in both ammonia-oxidizing archaea (AOA) and AOB. In the second step, oxidation of NH2OH to nitrite (NO2) occurs, catalyzed by hydroxylamine oxidoreductase (HAO). In the final step, nitrite-oxidizing bacteria containing the enzyme nitrite oxidoreductase (NIX) transform NO2 to NO3 [86,88,89]. It was recently shown that a group of bacterial species, i.e., comammox Nitrospira, can accomplish NH3 oxidation to NO3 within the same bacterial cell [89,90].
The reverse process of denitrification, where NO3 is reduced to NO2, NO, N2O, and N2, is also catalyzed by a diverse cast of bacterial and fungal species [86,91] (Equation (2)).
N O 3 nitrate reductase ( NR ) N O 2 nitrite reductase ( NIR ) N O nitric   oxide reductase ( NOR ) N 2 O nitrous   oxide reductase ( NOS ) N 2
The latter three products are gases, which are lost to the atmosphere. Of particular concern is the greenhouse gas N2O, which has a 273 times higher global warming potential than carbon dioxide (CO2) [92].
Attempts to fully understand and mitigate agricultural Nr losses have become the focus of intensive research initiatives. Numerous strategies have been put forward, such as coordinating fertilizer application methods with plant growth patterns, improving cropland management practices, and the use of synthetic urease and nitrification inhibitors [93,94]. The latter are typically applied as a formulation with the respective fertilizer.

4. Use of Inhibitors

Urease inhibitors (UIs) are synthetic chemicals that can be added to the urea-based fertilizer to slow down the catalytic hydrolysis of urea to NH4+ by preventing its binding to the enzyme urease (Figure 2). Urea is highly soluble in water and is often applied in agricultural systems through top-dressing. By slowing down its hydrolytic degradation, urea can travel deeper into the soil after rainfall or watering so that less NH3 escapes from the soil, leaving more nitrogen available for plant uptake through the growth cycle [95,96]. Furthermore, a slowed-down urea hydrolysis leads to a gradual NH4+ production (rather than one big burst), which enables a more efficient NH4+ uptake by plants, thereby reducing the extent of nitrification and the undesired loss of NO3 through leaching into the ground and surface water [96].
While many compound classes with the potential to inhibit urease activity in soils are known, phosphoric and thiophosphoric triamides, which can be considered as urea analogs, are the most promising synthetic urease inhibitors so far (Figure 2). In particular, N-(n-Butyl) thiophosphoric triamide (NBPT, trade name Agrotain) is effective in a wide range of soils at low concentrations of 0.01% of applied nitrogen [97]. The active inhibitor species is believed to be N-(n-butyl) phosphoric triamide (NBPTO), which is a urea analog that results in the hydrolysis of NBPT by soil microorganisms (Figure 3). The reason why in practice NBPT—and not NBPTO—is applied onto soils is due to the latter’s high susceptibility to hydrolysis, whereas the less nucleophilic P=S moiety increases the shelf-life, making the handling of this compound easier. A mixture of NBPT with the homolog N-(n-propyl) thiophosphoric triamide (NPPT) is marketed under the trade name Limus (BASF) [98]. Recently, NBPT has been co-formulated with Duromide, an NBPT derivative containing a covalently linked urea moiety (trade name ANVOL, Koch, Wichita, KS, USA), which increases NBPT efficiency by reducing NH3 volatilization.
Similarly, nitrification inhibitors (NIs) are intended to increase the residence time of NH4+ by delaying its oxidation to NO3 (Equation (1)) by inhibiting the AMO enzyme of nitrifying microbes in the soil responsible for this process. Many of the compounds with known nitrification-inhibitory properties are aromatic heterocycles containing two or more N, O, or S atoms, suggesting that NIs target the first step of the ammonia oxidation process, i.e., NH4+ → NH2OH, that is catalyzed by AMO. The second nitrification step, where NH2OH is oxidized to NO2 catalyzed by HOA, is normally not influenced [99].
It should be noted that plant roots can secrete compounds that can also slow down the rate of NH4+ oxidation by blocking both the AMO and HAO-catalyzed steps of the nitrification cascade [100,101,102]. For example, a study on Brachiaria pasture [101] led to the discovery of brachialactone as a new biological nitrification inhibitor (BNI), which contributed up to 90% of the inhibitory activity released from the roots of this tropical grass. Brachialactone blocks both AMO and HOA enzymatic pathways in Nitrosomonas [101], and within three years of establishment, Brachiaria pastures have suppressed soil nitrifiers, such as archaeal and bacterial ammonia-oxidizing microorganisms. A similar decrease was reported for two forage crops, Medicago sativa and Dactylis glomerata [103]. It was suggested that the release of BNIs is triggered by high NH4+ concentrations in the root environment [101,102]. This proposal was supported by a study, where half of the root system was exposed to NH4+ and the other half to NO3. Only the part exposed to NH4+ triggered the release of brachialactone, indicating a localized release process. Brachialactone is a tetracyclic diterpene with a unique 5-8-5-membered ring system, a γ-lactone ring, and seven stereogenic centers.
From the highly diverse (and complex) frameworks of these compounds, different modes of action could be suggested, although mechanistic details are not yet known. There is some controversy in the literature about whether the use of BNIs could be a viable strategy to modulate nitrification processes in agricultural soils [104]. From the viewpoint of synthetic organic chemistry, access to these BNIs would require multi-step processes, which is not feasible for the large-scale synthesis required for applications in broad-acre agricultural systems. On the other hand, genetic engineering approaches [105] might potentially provide an opportunity to create plants with the ability to excrete higher concentrations of NBIs. In this overview, we will focus only on synthetic small-molecule inhibitors and refer the reader interested in BNI to explore recent reviews in [5,106,107,108].
2-Chloro-6-(trichloromethyl)-pyridine (Nitrapyrin or N-Serve®, Dow Chemical Co., Midland, MI, USA) is the first synthetic nitrification inhibitor that has been brought to the market [109]. Since then, several compounds have been developed to control nitrification in soil (see Figure 2). The most important examples are dicyandiamide (DCD), 2-amino-4-chloro-6-methylpyrimidine, and 3,4-dimethylpyrazole phosphate (DMPP). Amongst these, the properties and performance of DCD and DMPP have been most widely studied [110]. The active site in DMPP is 3,4-dimethyl-1H-pyrazole phosphate (DMP), but because of its volatility, this inhibitor is usually applied as a salt with phosphoric acid.

4.1. Mechanism of Urease Inhibition (UI)

The mechanism by which NBPT (or NBPTO) inhibits urease activity can be understood from the interactions between urea or inhibitor, respectively, and the enzyme’s active site. Figure 4 shows that urea coordinates with the two Ni centers in the enzyme’s active site via the carbonyl oxygen and one of the amino groups, which activates the carbonyl group for nucleophilic attack by the OH bridge between the two Ni atoms. NBPTO is a tridentate ligand that forms a complex with both Ni atoms and the oxygen atom of the urea-derived carbamate, which mimics the transition state of urea hydrolysis. The lateral N-(n-butyl) amine substituent blocks access of urea to the active site. Unfortunately, in the NBPTO-urease complex, NBPTO is also activated for hydrolysis, which leads to the loss of the n-butyl amine substituent, which results in the deactivation of the inhibitor [111].
While NBPT can delay urea hydrolysis for 3–14 days, depending on soil and environmental conditions [112], the impact on the N2O emissions is highly variable, ranging from no effect to reduced or even increased N2O emissions [113]. Likewise, the benefits of NBPT for crop yield, N-uptake, and NUE are quite variable and span from some increase to no significant effect [114].
Many other compound classes with urease-inhibiting properties are known, which act through different mechanisms [115,116], for example, by preventing urea access through targeted “space-filling” of the active site involving noncovalent interactions with neighboring amino acid residues, typically histidine, aspartate, and lysine [117] (Figure 5a). Through analysis of the crystal structure of the active site and molecular docking studies, inhibitor compounds can be designed for specific applications [118].
Inhibition can also be achieved through modification of the cysteine-rich mobile flap, which allows access to the active site, through covalent [83] (and refer to Figure 5b (bottom)) or π-interactions [121]. Data from biochemical and mechanistic studies show the formation of a covalent C–S bond with 1,4-benzoquinone, a known inhibitor, which reduces the flexibility of the mobile flap and slows down or even prevents access of urea to the active site of the enzyme (Figure 5b). However, in the context of inhibition of urease in soils, the behavior of new inhibitor compounds in the complex soil matrix also needs to be considered: Mobility, as defined by molecule size, lipophilicity, hydrophilicity, and charge, as well as stability towards degradation are important factors [122].

4.2. Mechanism of Nitrification Inhibition (NI)

The biochemical mechanisms of nitrification inhibition and the effects of nitrification inhibitors on reducing nitrification rates have not been confirmed or verified [109]. However, based on the inhibitory effects of several compounds, it is suggested that NIs target the first step of the nitrification process, specifically the metalloenzyme enzyme AMO [123]. AMO contains copper(II) (Cu(II)) in its active site, and the essential role of Cu(II) in the catalytic process was discovered already in the 1940s [124]. However, the mechanism by which commercial inhibitors, for example, DMPP, inhibit AMO, is still unclear, which is likely due to the lack of knowledge of the structure of the active site in AMO. It has been recently revealed that DMP and DCD act as reversible NI and do not compete for the same binding site as NH3 [125], confirming the potential chelator properties of DMP and DCD. As a membrane-bound enzyme, its activity decreases significantly after isolation [126], hampering direct enzymatic studies so far. The discovery of the evolutionary similarity between AMO and the particulate methane monooxygenase (pMMO) has been an important milestone in gaining a better understanding of AMO, which emphasized the significant role of copper in its activity [127]. As an alternative to enzymic studies, strategies involving pure bacterial cultures have been employed to screen potential nitrification inhibitors [125]. Most commonly, Nitrosomonas europaea (NE) has been used as the model organism because it can be grown in a laboratory setup in high cell numbers [128]. The nitrification rate can be determined by a colorimetric assay (Griess Assay), where NO2 production is monitored over time [129]. More recently, archaea have been increasingly utilized in these assays as they are believed to be the more predominant species in soils [129,130,131,132]
The chemical transformations mediated by AMO have been explored with small molecules, including linear and cyclic alkanes, alkenes, aromatic compounds, as well as NH3 (Figure 6) [133,134]. Thus, AMO in NE catalyzes the insertion of oxygen into C-H bonds, enabling the conversion of alkanes or aromatic compounds to the corresponding alcohols. In the reaction with alkenes, the formation of epoxides occurs, whereas alcohols can be further oxidized to carbonyl compounds. NH3 is oxidized to NH2OH through the insertion of O into an N-H bond.
Importantly, the growth of nitrifying bacteria can only be maintained when NH3 is present [142]. This finding can be understood by examining the flow of electrons during the first two nitrification steps (Figure 7). Thus, the AMO-catalyzed oxidation of NH3 (or NH4+, respectively) to NH2OH releases two electrons, which are taken up by O2. The reduction O2 → 2 O2− requires four electrons. The other two electrons are provided by HAO, which supplies a total of four electrons through the oxidation NH2OH → NO2. The remaining two electrons are transferred to Cytochrome C554, membrane cytochrome Cm552, and Ubiquinone-8 (QH2). HAO is, therefore, a crucial enzyme in the nitrification process that fuels the AMO reaction [143,144]. In other words, if the compound produced by AMO is not a substrate for HAO, the entire oxidation cascade cannot be energetically maintained.
Based on this, AMO inhibition could be achieved via two pathways: (i) by substrate inhibition, where the inhibitor binds to AMO but its oxidation product is not a substrate for HAO and (ii) by ligand binding to Cu(II) in the enzyme’s active site, which maintains its catalytic function but prevents binding of NH3. The crystal structure of the N-terminus of the amoB subunit in AMO shows that the periplasmatic cupredoxin-like domain is formed by three histidine residues, which coordinate with Cu(II) [126,145]. These data suggest that the existing commercial NIs bind to Cu(II) in a similar fashion, thereby slowing down nitrification rates. This proposition is supported by crystallographic analyses of model systems, which revealed that four DMP molecules and two Cl ions can form an octahedral complex with Cu(II) [146,147], demonstrating the ability of DMP to act as a ligand for Cu(II). The findings from Corrochano-Monsalve et al. [146] further show that the growth of N2O-reducing bacteria appears to be controlled by Cu availability. These bacteria benefit from the use of DMPs, which could be attributed to a decrease in competence when AOB growth is inhibited. This opened up the possibility to induce N2O reduction to N2 through Cu fertilization.
Recent innovations in the development of new NI compounds are mainly based on the modification of existing inhibitors to increase their effectiveness. The most recent modification for DMPP is 3,4-dimethylpyrazole succinic acid (DMPSA), showing (reduces N2O flux rates) in trial field experiments [148]. DMPSA is believed to be stable in basic conditions and can be formulated with calcium ammonium nitrate [149] but is not (yet) commercially available. Another approach is slow-release inhibitors, which are intended to inhibit nitrification over a longer time. For example, a recent soil incubation study where the amount of volatilized NH3 was measured revealed that a copolymer of DMPP with acrylic acid (AA) can slow down nitrification over a duration of 24 h [131,150]. In this regard, it should be noted that many approaches to control the release of fertilizers (with or without inhibitors) have been made. This concept is outside the scope of this overview, and the interested reader is referred to a recent review by Vejan et al. [151].

4.3. Factors Affecting the Effectiveness of UIs and NIs

Temperature, soil pH, clay content, organic matter, and water-filled pore space (WFPS) are among the most important factors influencing the effectiveness of urease and nitrification inhibitors [152,153,154,155,156,157,158].

4.3.1. Temperature

A soil incubation study by Carmona et al. [159] found an increase in NH4+ production by 50% at 32 °C than at 18 °C, suggesting that higher application rates of NBPT would be required at higher temperatures to enable inhibition of urea hydrolysis to a similar extent than at cooler temperatures. Suter et al. [113] showed that 55% of urea was still present in the soil at temperatures between 5 and 15 °C two weeks after NBPT/urea treatment, while only <2% of urea remained at 25 °C. The decreased inhibitory efficacy at higher temperatures is likely due to an increased rate of hydrolytic degradation of NBPT, which can become a significant factor in the field during periods of temperatures > 35 °C that can occur during summer in certain places of the world. An additional challenge to inhibitor activity arises from the fact that the activity of urease also increases with increasing temperatures [160]. However, it should be noted that limited studies have specifically investigated the effects of lower soil temperatures on the performance of NBPT, indicating a need for further research in this area.
A decreased effectiveness of nitrification inhibitors with increasing temperature has also been found. A field study by McGeough et al. [161] revealed that the measured half-lives (t1/2) of DCD across different mineral soils decreased with increasing soil temperature, suggesting that the efficacy of DCD depends not only on soil temperature but also on soil composition, such as clay content and organic matter. Kelliher et al. [162] showed in field trails that t1/2 of DCD was approximately half of that measured at the same temperature under laboratory conditions, whereas the relationship between t½ mean soil temperature is: (t½ (days) = 54 − 1.8 T (°C). Similarly, the inhibitory activity of DMPP is also inversely related to temperature, with a significant drop in activity observed over a relatively small temperature change [163]. For example, it has been shown that at a soil temperature of 35 °C, which can occur during fertilization application in some South Asian regions, DMPP remains effective for only one week [164]. Using DMPP to mitigate N2O emissions from sheep urine patches revealed a poorer efficiency over summer [165].

4.3.2. Soil pH

Deprotonation of NH4+ readily occurs in alkaline soils, which could lead to higher NH3 volatilization. Thus, the contribution of urease inhibitors to reducing such unwanted NH3 losses is relatively higher in alkaline than in neutral or acidic soils [156]. As mentioned previously, NBPT and NBPTO are both prone to hydrolytic degradation, in particular at low pH (<5) [166,167], whereas even at pH ~ 7, abiotic chemical hydrolysis of NBPT still occurs to some extent. Thus, a reduction in urea hydrolysis of about 17–86% by NBPT was reported by Tao et al. [168] in acidic soils compared to a 53–92% reduction in alkaline soil [166,167]. At higher pH (>8), microbial activity is likely responsible for the observed increase in the degradation rate of NBPT [167,168], as shown by the higher t½ of NBPT in sterilized compared to nonsterilized soil [165]. The significance of microbial degradation is further highlighted by a slightly longer t½ of NBPT in light compared to dark conditions (5.9 vs. 4.5 days, respectively), which could be attributed to the light-inhibiting microbial growth and therefore a slower degradation of NBPT [168].
Similarly, several studies [169,170,171,172,173] have shown that the distribution and abundance of ammonia oxidizers and therefore the efficacy of nitrification inhibitors are influenced by soil pH. The generally poor inhibitory performance of DMPP at low pH could result from the switch from the autotrophic AOB, which are targeted by DMPP, to AOA, which are not inhibited by DMPP [87,174]. AOA have a lower metabolic concentration threshold to using NH3 and contribute to the nitrification process more than AOB in very low pH soils, where NH3 availability is low [175,176]. While the mechanism of DMPP inhibition is not well understood [125], it has been shown that DMPP decreases the abundance and metabolic activity of AOB in alkaline and, to a lesser extent, in acidic soils (due to the lower AOB populations in these soils) but does not have an inhibitory effect on the AOA population in either soil type [177].

4.3.3. Clay Content and Organic Matter

The inhibitory performance of NBPT and other urease inhibitors is also affected by soil organic matter (OM), where higher OM content results in increased adsorption and reduced effectiveness due to lower mobility [178]. Furthermore, in soils with a higher OM, the microbial activity is also increased, resulting in faster inhibitor degradation and therefore reduced efficacy. Christianson et al. [179] reported that NBPT changed soil pH and NH4+ concentration around the reaction site in loamy silt soil with 20% clay content, resulting in the diffusion of urea from the placement site. On the other hand, at a site with 46% clay soil, the inhibitory activity of NBPT was low, as revealed by an increased NH4+ concentration around the fertilized site.
Similarly, McGeough et al. [161], Zhang et al. [180], Wu et al. [181], and Barth et al. [182] reported that the specific composition of organic matter and clays affects DCD and DMPP efficacy. The lower recovery rates of DMPP from clay soil have been rationalized to be caused by greater interaction with the organic matter [177]. Another study by Chen et al. [183] was performed on brown Vertosol soil with clay loam texture under varying moisture and temperature conditions, where it was shown that DMPP reduced cumulative N2O emission over 42 days by more than 65% in clay loam soil.

4.4. Use of Combined UIs and NIs

In principle, by using both urease and nitrification inhibitors together, increased N use efficiency should be achieved, as these inhibitors target the major N loss mechanisms through NH3 volatilization, NO3- leaching, and N2O emissions. However, although the NUE of crops can be increased by minimizing N losses, minimizing losses through one pathway could increase losses through the other [184]. For example, it has been shown that when NBPT is used alone, urea hydrolysis is slowed down, but once NH4+ is produced, it is exposed to nitrification and denitrification losses [153]. In particular, increases in N2O emissions with the use of UIs have been reported [185,186,187]. Similarly, when NIs are used alone, the prolonged retention of NH4+ in the soil can increase the extent of NH3 volatilization [186,187,188,189,190,191], indicating that the use of NIs may not be beneficial in areas with concerns about NH3 emissions. In fact, some NIs were found to increase NH3 loss [23,24,188,192,193]. Thus, an approach that simultaneously targets all N pathways for an overall environmental benefit is required. Dual-action inhibitors could potentially provide a solution, where UIs, NIs, and ammoniacal (NH3-N) fertilizers are applied in combination to minimize losses from the N-cycling pathways of hydrolysis, nitrification, and denitrification [189,192,194].
However, while such an approach seems to be “logical”, several studies have revealed that the combination of both inhibitors with urea does not necessarily give the expected beneficial effect. For example, a soil incubation study where urea was applied together with NBPT and DCD revealed that DCD could even offset the effect of NBPT in reducing the loss of NH3 through volatilization (Figure 8) [191]. As the inactivation of NBPT by DCD could be excluded (and vice versa), these data suggest that the slowed-down nitrification process leads to the accumulation of NH3/NH4+ in the soil, allowing volatilization losses to continue.

5. Use of Nitrification and Urease Inhibitors in Different Plant Genotypes—What Is Known?

Slowing down urea hydrolysis in the soil through the use of UIs is also beneficial for crops, as it reduces the toxic effect of NH3/NH4+ concentration spikes in the area surrounding fertilizer granules on seed germination, resulting in higher crop yields (barley and canola) and increased root growth (rice) [195,196,197].
However, NBPT may be taken up by plant roots, as has been demonstrated in maize seedlings through a reduced net urea uptake and lower ureic-N accumulation [198]. NBPT absorption by plants can lead to changes in metabolic pathways associated with N assimilation, such as a reduction in plant urease and glutamine synthetase [199,200]. A study by Cruchaga et al. [201] revealed that NBPT absorbed by pea plants caused inhibition of urease activity in the leaves and roots, which resulted in urea build-up and therefore necrosis of the plant at leaf margins. Other studies also found that NBPT applied at high concentrations of urea (0.5% w/w) and concentrations of 10 μg/g of soil was associated with leaf tip scorch. This effect is likely dose-dependent [202,203], with the leaf tip scorch believed to result from a toxic build-up of urea at the leaf tip rather than a cytotoxic effect exerted by NBPT itself. Cruchaga et al. [201] also found that NBPT taken up by pea and spinach roots is translocated to the leaves, potentially inhibiting the activity of endogenous leaf and root urease. Thus, NBPT can cause transient yellowing of leaf tips caused by urea toxicity soon after application. However, it has been reported that plants recover quickly and no effects on growth have been reported [199,204].
Similar to urease inhibitors, nitrification inhibitors added to N fertilizers can temporarily reduce the populations of nitrifying bacteria in the soil which are responsible for converting NH3 to NO2 (Nitrosomonas) and NO2 to NO3 (Nitrobacter). This process is important to plants as it leaves an extra stash of NH4+ that can be absorbed by the plants through their root systems. Few studies have reported the uptake of nitrification inhibitors by plants and their movements in plant tissues. The Kow (a relative indicator of an organic compound’s ability to bind to soil) value of −1 for the nitrification inhibitor DCD shows that this is a polar compound, which should be able to translocate in the trans-vascular system of plants. Likewise, it has been found that plants have the capacity to take up DMPP and DMPSA [205]. This study was conducted on clover plants, which revealed that DMPP (applied at 100 mg/kg soil) accumulated in clover leaves where it caused necrosis on the leaf margins. On the other hand, DMPSA (applied at 100 mg/kg soil) accumulated in the roots with no harm to the root tissue. As the application rates in this study were significantly higher than the estimated maximum amount used in agriculture (0.5 mg/kg soil), both inhibitors pose an extremely low risk of phytotoxicity [205]. Similarly, some studies are available that identified Nitrapyrin residues in food crops, providing evidence that plants also take up and accumulate this nitrification inhibitor. For example, a gas–liquid chromatographic study under field conditions by Kallio et al. [206] revealed that the amount of Nitrapyrin residues in red beetroots was strongly correlated with the application rate of this inhibitor.
Recently, in two separate field studies, it was shown that urea with Nitrapyrin alone or in combination with the potassium salt of gibberellic acid (GA-K salt) has the potential to enhance NUEs and yields of maize and wheat [207,208]. Thus, it was found that maize plant biomass, grain yield, and total N uptake were highest and increased significantly by 27%, 36%, and 25%, respectively, in the treatment with urea, Nitrapyrin, and GA-K salt compared to urea treatment alone or urea with Nitrapryrin. Similarly, wheat plant biomass, grain yield, and total N uptake were enhanced by 31, 37, and 44%, respectively, when urea was applied with Nitrapyrin and GA-K salt as compared to urea alone or used in combination with Nitrapyrin.
Meta-analyses have also been widely performed on various crops to study the effect of a series of inhibitors on plant productivity. Linquist et al. [209] explored the effect of several N fertilizers amended with either NIs or UIs on yield and N uptake in rice systems. The study found that, on average, the use of these fertilizers led to a 5.7% increase in yield and an 0.8% increase in N uptake. Abalos et al. [210] integrated available results to quantitatively evaluate the effect of commonly used NIs and UIs on crop productivity. A meta-analysis was conducted to characterize the response of crop productivity with the application of nitrification (DCD, DMPP), urease (NBPT), and combined use of both inhibitors (DCD + NBPT). This study demonstrated that co-application of NIs or UIs is generally higher under conditions that favor high drainage and if high inputs of fertilizer were applied [210]. Another meta-analysis by Yang et al. [211] compared the effectiveness of DCD and DMPP across maize field trials, revealing that both inhibitors were equally effective in altering soil inorganic N content, reducing inorganic N leaching and nitrous oxide (N2O) emissions. However, DCD was more effective than DMPP in increasing plant productivity (6% increase in grain yields).

6. Future Outlook/Perspectives

Global cycling of N has reached its peak—the amount of atmospheric N2 industrially converted into NH3 to produce fertilizers now exceeds the amount that is cycled through all of the Earth’s terrestrial processes [15]. One of the strategies to increase N-use efficiency in agricultural systems is through the use of enzyme inhibitors that modulate the microbial conversion of fertilizer N in soils. Thus, the hydrolysis of urea to NH4+ by the enzyme urease present in soils can be delayed by the use of urease inhibitors. Likewise, the oxidation of NH3 to NO3 and its accumulation during the cropping phase can be regulated by nitrification inhibitors. Urease inhibitors have been on the market for more than 20 years, with growing acceptance by farmers [192]. While NBPT, currently the most important commercial UI, can reduce (but not eliminate) NH3 volatilization [192], serious limitations, such as the short period of effective inhibition in conjunction with a limited shelf life, should stimulate much-needed fundamental research towards the development of improved UIs that delay hydrolysis of surface-applied urea until the full dissolution in the soil body. Similarly, several nitrification inhibitors are currently on the market, such as Nitrapyrin, DMPP, and DCD, which slow down NH3 oxidation and reduce N loss through NO3 leaching and denitrification. However, there is inconsistency in both environmental and agronomic benefits of the use of inhibitors due to the complex interactions between soil, crop, and agricultural management in conjunction with climate (and climate changes) occurring in cropping systems.
This review provides an evaluation of the current knowledge of the available inhibitor technology and its shortcomings. This assessment reveals that the development of novel inhibitors with consistent performance across many different agricultural settings requires an improved mechanistic understanding of the spatial and temporal interplay of urease and nitrification inhibitors with plants and N uptake by roots.
Various approaches have been suggested to increase the use efficiency of nitrogenous fertilizers with inhibitors, such as (i) improved agricultural practices, for example, increased frequency of crop rotation or farmers implementing best management practices with regards to the 4 Rs Nutrient Stewardship—the “right” rate, source, timing, and placement of nitrogen fertilizers to match plant needs; (ii) using coatings, which include rapid, low-cost, ideally one-step coating methods for fertilizers, for example, metal–ligand coordination complexes or biodegradable/compostable polymers that enable nutrient release upon environmental triggers, for example, a rainfall or pH changes [212,213]; or (iii) reducing NH3 volatilization arising from urea hydrolysis by using inexpensive additives which are readily available as natural minerals or as industrial by-products, and through chemical and physical modifications. For example, zeolites or surface-modified coal tailings can be used to reversibly capture NH3 [214], which could reduce the risk of N loss through runoff or denitrification. However, these approaches are only useful if the crop can indeed take up and use the saved N.
The benefits of using inhibitors to mitigate N losses have been explored in several field trials. For example, Meng et al. [215] investigated the effect of DMPP and DCD on N2O emissions and inorganic-N losses through leaching and runoff in a mixed pasture field. Although this study explored only one inhibitor concentration, positive effects of these NIs on greenhouse gas reduction, water quality, and improved soil health were found. Another field trial compared DCD and DMPP on altering soil inorganic N content in two soils with different pH and found that DCD is more effective than DMPP in increasing plant productivity [216]. However, one general disadvantage of these experimental approaches has been that only one variable was evaluated to test the efficiency of the inhibitors, such as soil pH, different inhibitor concentrations, or different crops. It is known that interactions among these variables can have a significant impact on system responses and therefore studying components in isolation is not necessarily sufficient to draw conclusions about the entire cropping system.
The development of new inhibitors requires a collaborative approach involving synthetic chemists, biochemists, microbiologists, plant nutritionists, and agronomists. Thus, to understand the overall benefit of the existing and newly developed inhibitor compounds, studies addressing their interactions with the soil components, in particular, the targeted enzymes in the soil (i.e., urease and AMO) and the root–soil system in dependence on concentration as well as their fate in the soil (i.e., identification of degradation products and degradation mechanism) in the absence and presence of plant roots are required. Factors such as microbial population, organic matter, and clay content are some examples of other variables that could affect inhibitor performance and degradation in soil. In other words, a better understanding of the chemical and microbiological processes that are triggered when an inhibitor is added to the soil is urgently required. For example, some known urease inhibitors, such as benzohydroxamic acid, perform well within in vitro assay experiments but poorly in soil incubation tests, possibly due to soil–metal ion interaction. This poses a problem as many urease inhibitors aim to chelate to the nickel centers within the active site and cannot perform inhibition if they are trapped by metal ions present in the soil. Therefore, understanding more about the fate of inhibitors and the effect of various soil variables by analyzing the inhibitor concentration–time profile in soils and their degradation products is essential for the design of new inhibitors with reliable performance.
This research could be further assisted by simulations using predictive models to create an artificial biological system, which will allow to link soil properties and biomass production in silico [217]. For example, dynamic system simulation models (DSSMs) describe the changes in system states in response to external drivers, such as management practices, weather, and soil properties, and how these changes are affected by other components in the system. This approach can be used for all types of crop models and farming system models [217].
The development of novel inhibitor compounds will benefit from approaches used in the pharmaceutical drug discovery process, in particular, structure–activity relationship (SAR) studies, which assess the performance of different compounds in dependence on their molecular structure. SAR studies require large libraries of compounds, with their synthesis often becoming the rate-limiting step. One way to facilitate this design process is through using artificial intelligence (AI) [218]—a “deep learning” computer program that produces blueprints for the sequences of reactions needed to create new inhibitors. Pharmaceutical companies have been using AI to streamline the process of discovering new medicines for years [219]. For example, Segler and his team [220] used deep-learning neural networks to imbibe nearly all known single-step organic-chemistry reactions—approximately 12.4 million of them. The AI tool employs these neural networks repeatedly in the planning of a multi-step synthesis, deconstructing the desired molecule until it reaches the available starting reagents. The program-generated reaction pathways were validated in a double-blind trial, where the quality of the machine-devised synthetic route was compared to that designed by synthetic chemists. This example shows that the use of AI could be a promising approach to streamline the development of new inhibitor compounds.
Given the scale of production required for use in agriculture, new inhibitor compounds should be accessible in a few steps from readily available precursors. Furthermore, they need to be safe in handling, nontoxic to other soil organisms, aquatic life, crops and safe to the environment, efficient, and economical in use, as well as chemically stable over the duration of the application and decomposing to environmentally unproblematic products. The SAR-guided design of inhibitors based on the enzyme’s active site is a useful approach for the design of urease inhibitors, as the active site has been widely studied in many species and remains largely conserved among them. Detailed crystallographic data are readily available that enable docking studies to assess binding modes and affinity of substrate–ligand complexes to guide inhibitor design before embarking on the synthesis [97,221]. Successful inhibitors later in the developmental process may be able to be co-crystallized with urease to obtain an understanding of the mode of binding. Additionally, prior to lengthy soil incubation experiments, inhibitor candidates can be screened for potential urease inhibition properties through an in vitro assay with Jack-Bean urease. Through this assay, NH3 production can be quantified using a colorimetric technique developed by Weatherburn [222]. This approach is quick, reproducible, and allows for the rapid screening of a large library for inhibitory activity.
In the case of AMO, similar inhibitor–enzyme studies are hampered due to the lack of structural information on the active center. In this case, SAR studies could provide novel information about its structure. Rather than performing time-consuming soil incubations as the first performance screen, whole cell assays using commercially available AOB can be used to determine biochemical parameters in the absence of soil. For instance, a quick 60 min nitrification assay can determine the % inhibition and IC50 value of new compounds in vitro. Once compounds with promising inhibitory effects have been identified more complex studies should be performed. One important tool is real-time kinetic studies, with a Clark-type electrode to distinguish between irreversible from reversible inhibitors of AMO [125,134]. In this experiment, the oxygen respiration rate of AOB is measured in situ prior to and after the addition of the NI to monitor the real-time kinetic order of the reaction [125,223]. Additionally, in-depth binding studies, such as Michaelis–Menten kinetics, reversibility of binding, and relative toxicity, are substantial experiments that can be incorporated into the pipeline to the development of new NIs [125,131].
Following the enzymatic studies using the JBU assay for UIs or the bacterial assay for NIs, promising candidates will be further evaluated in soil incubation studies without and with plants before moving to field trials.
A major consideration for the selection of UIs and NIs is their high effectiveness at the lowest possible application rate with a minimum of undesirable side effects. Uniform interaction of the substrate with the inhibitor can be achieved by either coating fertilizer granules with the inhibitor or by incorporating it into granules, as suggested by Slangen and Kerkhoff [224]. However, optimizing the inhibitor/N-fertilizer application rate is a complex task since it involves consideration of several variables, for example, different crops and cropping systems, soil properties, water availability, and climate. Using a bottom-up approach, where the complexity of the system is successively increased, could be the best strategy to address these problems. Plant–root interactions could be explored using simpler plant growth systems, such as Root-TRAPR [225] and EcoFAB [226]. These systems facilitate the exploration of root morphology and root exudate under controlled conditions and at a smaller scale. The devices are easy to build and can be used to study plant responses to elicitor challenges. Furthermore, to measure and validate the agronomic and environmental benefits of the new fertilizer formulations (with NIs and UIs), glasshouse and field trials for a range of production systems including vegetables and dairy pasture need to be carried out. This will help to evaluate the nitrogen loss pathways and yield benefits of the newly developed compounds. Measurements of N2O emissions and evaluation of potential ecotoxic effects will provide insight into the environmental impact of these new compounds. It is apparent that this program requires the cooperation of researchers from various disciplines.
One of the considerations for field trials could be the method of partial ammonium nutrition provision to crops. This means that ammonium can be a direct source of nitrogen for crops instead of nitrate. It is poorly translocated to the rhizosphere, unlike nitrate, which prevents its rapid uptake by plant root–soil systems. Ammonium nutrition to crops also improves the uptake of phosphorus. Thus, protons (H+) are excreted to maintain charge equilibration in the roots when plants’ roots take up NH4+, the resulting pH decrease in the rhizosphere supports the mobilization of phosphorus in the soil [227] and also increases the uptake of some micronutrients, such as manganese (Mn) [228]. Thus, by increasing the residence time of NH4+ in soil by applying a nitrification inhibitor, the effect on phosphate and micronutrients mobilization for plant nutrition can be intensified.
Another strategy to optimize N inputs is using the split application of N fertilizers in the soil. The split application involves applying smaller amounts of fertilizer at various times throughout a crop’s growth cycle [229] (Halvorson et al., 2008). Rozas et al. [230] and Burton et al. [231] found that N2O emissions were reduced by delaying part of the application of N fertilizer from irrigated maize and in a warm wet year, respectively. In another study by Zhang et al. [232], it was revealed that the use of split nitrogen fertilizer application improved the antioxidant enzyme activity and the remobilization of photosynthate after anthesis in crops, thereby improving wheat grain yield. The use of nitrification and urease inhibitors along with the split application method could further reduce NO3 concentrations, thereby reducing leaching and N2O emissions, providing a greater yield, high-quality characteristics, and balanced nitrogen management [233].

7. Conclusions

In conclusion, optimizing N availability to plants is crucial for maximizing crop yield and quality. However, the challenge lies in managing the various pathways through which nitrogen can be lost, including volatilization, emissions, leaching, runoff, and immobilization. While synchronizing N availability with plant uptake could potentially reduce N loss, it is influenced by both controllable management practices and uncontrollable weather dynamics. In recent years, the use of urease and nitrification inhibitors has emerged as a promising strategy to mitigate nitrogen loss and synchronize N availability with plant uptake. Urease inhibitors delay the hydrolysis of urea to NH4+, thereby reducing nitrogen loss through ammonia volatilization. Nitrification inhibitors temporarily inhibit the conversion of NH4+ to NO2 by soil bacteria, slowing down the nitrification process and minimizing nitrogen loss as NO3 or through denitrification.
This review has provided a comprehensive understanding of urease and nitrification-inhibitor technologies and their profound implications for plant growth and root nitrogen uptake. It emphasizes the need for developing inhibitors with enhanced efficiency and design principles. These inhibitors have the potential to revolutionize agricultural practices and play a significant role in optimizing nitrogen management for sustainable crop production. Looking ahead, future directions for inhibitor usage should focus on harnessing innovative advancements that prioritize the essential traits superior inhibitors should possess. By doing so, we can pave the way for more efficient and sustainable nitrogen management in agriculture.
In summary, the utilization of urease and nitrification inhibitors offers a promising approach to optimize nitrogen availability to plants, reduce nitrogen loss, and enhance crop production. Continued research and development in this field will contribute to the advancement of sustainable agricultural practices, ensuring improved nitrogen management and meeting the growing global demand for food while minimizing environmental impacts.

Author Contributions

S.G. conceived the general idea of the manuscript. S.G., U.R. and U.W. developed the concept. S.G., S.Y. and B.A. wrote the manuscript with the help of U.R. and U.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Australian Research Council’s Discovery Grant DP200101162 and the Australian Research Council’s Industrial Transformation Research Program funding scheme IH200100023.

Data Availability Statement

No new data were created or analysed in this study. Data sharing is not applicable to this article.

Acknowledgments

We kindly acknowledge that S.G., U.W., and U.R. have received financial support from an Australian Research Council’s Discovery Grant DP200101162. S.G., S.Y., and B.A. are also part of ARC Research Hub for Smart Fertilisers which is under the Australian Research Council’s Industrial Transformation Research Program funding scheme IH200100023. S.Y. is thankful for financial support from a University of Melbourne Research Scholarship provided through the Jülich Melbourne Postgraduate Academy (JUMPA).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bernhard, A. The nitrogen cycle: Processes. Play. Hum. 2010, 3, 10–25. [Google Scholar]
  2. Erisman, J.W.; Galloway, J.N.; Seitzinger, S.; Bleeker, A.; Dise, N.B.; Roxana Petrescu, A.M.; Leach, A.M.; de Vries, W. Consequences of human modification of the global nitrogen cycle. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 2013, 368, 20130116. [Google Scholar] [CrossRef] [PubMed]
  3. Galloway, J.N.; Cowling, E.B. Reactive nitrogen and the world: 200 years of change. AMBIO 2002, 31, 64–71. [Google Scholar] [CrossRef] [PubMed]
  4. LeBauer, D.S.; Treseder, K.K. Nitrogen limitation of net primary productivity in terrestrial ecosystems is globally distributed. Ecology 2008, 89, 371–379. [Google Scholar] [CrossRef] [PubMed]
  5. Coskun, D.; Britto, D.T.; Shi, W.; Kronzucker, H.J. Nitrogen transformations in modern agriculture and the role of biological nitrification inhibition. Nat. Plants. 2017, 3, 17074. [Google Scholar] [CrossRef] [PubMed]
  6. Foley, J.A.; Ramankutty, N.; Brauman, K.A.; Cassidy, E.S.; Gerber, J.S.; Johnston, M.; Mueller, N.D.; O’Connell, C.; Ray, D.K.; West, P.C.; et al. Solutions for a cultivated planet. Nature 2011, 478, 337–342. [Google Scholar] [CrossRef]
  7. Rai, A.K. Agricultural nitrogen management for sustainable development and global food security. J. Sci. Res. 2019, 63, 79–87. [Google Scholar]
  8. Erisman, J.W.; Sutton, M.A.; Galloway, J.; Klimont, Z.; Winiwarter, W. How a century of ammonia synthesis changed the world. Nat. Geosci. 2008, 1, 636–639. [Google Scholar] [CrossRef]
  9. Bodirsky, B.L.; Popp, A.; Lotze-Campen, H.; Dietrich, J.P.; Rolinski, S.; Weindl, I.; Schmitz, C.; Muller, C.; Bonsch, M.; Humpenoder, F.; et al. Reactive nitrogen requirements to feed the world in 2050 and potential to mitigate nitrogen pollution. Nat. Comm. 2014, 5, 3858. [Google Scholar] [CrossRef]
  10. Senauer, B. Feeding the World: A Challenge for the Twenty-First Century; Oxford University Press: New York, NY, USA; Oxford, UK, 2001; Volume 83, pp. 790–792. [Google Scholar]
  11. Glibert, P.M.; Harrison, J.; Heil, C.; Seitzinger, S. Escalating worldwide use of urea—A global change contributing to coastal eutrophication. Biogeochemistry 2006, 77, 441–463. [Google Scholar] [CrossRef]
  12. Cassman, K.G.; Dobermann, A.; Walters, D.T. Agroecosystems, nitrogen-use efficiency, and nitrogen management. AMBIO 2002, 31, 132–140. [Google Scholar] [CrossRef] [PubMed]
  13. Ladha, J.K.; Pathak, H.; Krupnik, T.J.; Six, J.; van Kessel, C. Efficiency of fertilizer nitrogen in cereal production: Retrospects and prospects. Adv. Agron. 2005, 87, 85–156. [Google Scholar]
  14. Erisman, J.W.; Galloway, J.; Seitzinger, S.; Bleeker, A.; Butterbach-Bahl, K. Reactive nitrogen in the environment and its effect on climate change. Curr. Opin Environ. Sustain. 2011, 3, 281–290. [Google Scholar] [CrossRef]
  15. Fowler, D.; Coyle, M.; Skiba, U.; Sutton, M.A.; Cape, J.N.; Reis, S.; Sheppard, L.J.; Jenkins, A.; Grizzeti, B.; Galloway, J.N.; et al. The global nitrogen cycle in the twenty-first century. Philos. Trans. R. Soc. Lond. B. Biol. Sci. 2013, 368, 20130164. [Google Scholar] [CrossRef] [PubMed]
  16. Schlesinger, W.H. On the fate of anthropogenic nitrogen. Prod. Natl. Acad. Sci. USA 2009, 106, 203–208. [Google Scholar] [CrossRef]
  17. Behera, S.N.; Sharma, M. Investigating the potential role of ammonia in ion chemistry of fine particulate matter formation for an urban environment Science. Total Environ. 2010, 408, 3569–3575. [Google Scholar] [CrossRef]
  18. Gong, L.; Lewicki, R.; Griffin, R.J.; Tittel, F.K.; Lonsdale, C.R.; Stevens, R.G.; Pierce, J.R.; Malloy, Q.G.J.; Travis, S.A.; Bobmanuel, L.M.; et al. Role of atmospheric ammonia in particulate matter formation in Houston during summertime. Atmos. Environ. 2013, 77, 893–900. [Google Scholar] [CrossRef]
  19. Wang, S.; Nan, J.; Shi, C.; Fu, Q.; Gao, S.; Wang, D.; Cui, H.; Saiz-Lopez, A.; Zhou, B. Atmospheric ammonia and its impacts on regional air quality over the megacity of Shanghai, China. Sci. Rep. 2015, 5, 15842. [Google Scholar] [CrossRef]
  20. Makino, A. Photosynthesis, grain yield, and nitrogen utilization in rice and wheat. Plant Physiol. 2011, 155, 125–129. [Google Scholar] [CrossRef]
  21. Lin, Y.L.; Chao, Y.Y.; Huang, W.D.; Kao, C.H. Effect of nitrogen deficiency on antioxidant status and Cd toxicity in rice seedlings. Plant Growth Regul. 2011, 64, 263–273. [Google Scholar] [CrossRef]
  22. Luo, B.F.; Du, S.T.; Lu, K.X.; Liu, W.J.; Lin, X.Y.; Jin, C.W. Iron uptake system mediates nitrate-facilitated cadmium accumulation in tomato (Solanum lycopersicum) plants. J. Exp Bot. 2012, 63, 3127–3136. [Google Scholar] [CrossRef] [PubMed]
  23. Chapin, F.S.; Matson, P.A.; Mooney, H.A. Terrestrial nutrient cycling. In Principles of Terrestrial Ecosystem Ecology, 1st ed.; Springer: New York, NY, USA, 2002; pp. 97–122. [Google Scholar]
  24. Courty, P.E.; Smith, P.; Koegel, S.; Redecker, D.; Wipf, D. Inorganic nitrogen uptake and transport in beneficial plant root-microbe interactions. Crit. Rev. Plant Sci. 2015, 34, 4–16. [Google Scholar] [CrossRef]
  25. Jones, D.L.; Healey, J.R.; Willett, V.B.; Farrar, J.F.; Hodge, A. Dissolved organic nitrogen uptake by plants—An important N uptake pathway? Soil. Biol. Biochem. 2005, 37, 413–423. [Google Scholar] [CrossRef]
  26. von Wirén, N.; Gazzarrini, S.; Frommer, W.B. Regulation of mineral nitrogen uptake in plants. Plant Soil 1997, 196, 191–199. [Google Scholar] [CrossRef]
  27. Britto, D.T.; Kronzucker, H.J. NH4+ toxicity in higher plants: A critical review. J. Plant Physiol. 2002, 159, 567–584. [Google Scholar] [CrossRef]
  28. Cui, J.; Yu, C.; Qiao, N.; Xu, X.; Tian, Y.; Ouyang, H. Plant preference for NH4+ versus NO3 at different growth stages in an alpine agroecosystem. Field Crops Res. 2017, 201, 192–199. [Google Scholar] [CrossRef]
  29. Miller, A.E.; Bowman, W.D.; Suding, K.N. Plant uptake of inorganic and organic nitrogen: Neighbor identity matters. Ecology 2007, 88, 1832–1840. [Google Scholar] [CrossRef]
  30. Kuzyakov, Y.; Xu, X. Competition between roots and microorganisms for nitrogen: Mechanisms and ecological relevance. New Phytol. 2013, 198, 656–669. [Google Scholar] [CrossRef] [PubMed]
  31. Chen, W.-M.; Moulin, L.; Bontemps, C.; Vandamme, P.; Béna, G.; Boivin-Masson, C. Legume symbiotic nitrogen fixation by β-Proteobacteria is widespread in nature. J. Bacteriol. 2003, 185, 7266–7272. [Google Scholar] [CrossRef]
  32. Moulin, L.; Munive, A.; Dreyfus, B.; Boivin-Masson, C. Nodulation of legumes by members of the β-subclass of Proteobacteria. Nature 2001, 411, 948–950. [Google Scholar] [CrossRef]
  33. Graham, P.H.; Vance, C.P. Legumes: Importance and constraints to greater use. Plant Phys. 2003, 131, 872–877. [Google Scholar] [CrossRef] [PubMed]
  34. Mahmud, K.; Makaju, S.; Ibrahim, R.; Missaoui, A. Current progress in nitrogen fixing plants and microbiome research. Plants 2020, 9, 97. [Google Scholar] [CrossRef]
  35. Hachiya, T.; Sakakibara, H. Interactions between nitrate and ammonium in their uptake, allocation, assimilation, and signaling in plants. J. Exp. Bot. 2017, 68, 2501–2512. [Google Scholar] [CrossRef]
  36. Wang, M.Y.; Siddiqi, M.Y.; Ruth, T.J.; Glass, A.D.M. Ammonium uptake by rice roots. (I. Fluxes and subcellular distribution of 13NH4+). Plant Physiol. 1993, 103, 1249–1258. [Google Scholar] [CrossRef] [PubMed]
  37. Unkles, S.E.; Wang, R.; Wang, Y.; Glass, A.D.; Crawford, N.M.; Kinghorn, J.R. Nitrate reductase activity is required for nitrate uptake into fungal but not plant cells. J. Biol. Chem. 2004, 279, 28182–28186. [Google Scholar] [CrossRef]
  38. Tegeder, M.; Masclaux-Daubresse, C. Source and sink mechanisms of nitrogen transport and use. New Phytol. 2018, 217, 35–53. [Google Scholar] [CrossRef]
  39. Noguero, M.; Lacombe, B. Transporters involved in root nitrate uptake and sensing by Arabidopsis. Front. Plant. Sci. 2016, 7, 1391. [Google Scholar] [CrossRef] [PubMed]
  40. Léran, S.; Varala, K.; Boyer, J.C.; Chiurazzi, M.; Crawford, N.; Daniel-Vedele, F.; Lacombe, B. A unified nomenclature of nitrate transporter 1/peptide transporter family members in plants. Trends Plant Sci. 2014, 19, 5–9. [Google Scholar] [CrossRef]
  41. Cerezo, M.; Tillard, P.; Filleur, S.; Munos, S.; Daniel-Vedele, F.; Gojon, A. Major alterations of the regulationof root NO3 uptake are associated with the mutation of NRT21 and NRT22 genes in Arabidopsis. Plant Phys. 2001, 127, 262–271. [Google Scholar] [CrossRef]
  42. Filleur, S.; Dorbe, M.F.; Cerezo, M.; Orsel, M.; Granier, F.; Gojon, A.; Daniel-Vedele, F. An Arabidopsis T-DNA mutant affected in NRT2 genes is impaired in nitrate uptake. FEBS Lett. 2001, 489, 220–224. [Google Scholar] [CrossRef]
  43. Little, D.Y.; Rao, H.Y.; Oliva, S.; Daniel-Vedele, F.; Krapp, A.; Malamy, J.E. The putative high-affinity nitrate transporter NRT21 represses lateral root initiation in response to nutritional cues. Prod. Natl. Acad. Sci. USA 2005, 102, 13693–13698. [Google Scholar] [CrossRef]
  44. Orsel, M.; Chopin, F.; Leleu, O.; Smith, S.J.; Krapp, A.; Daniel-Vedele, F.; Miller, A.J. Characterization of a two-component high-affinity nitrate uptake system in Arabidopsis Physiology and protein-protein interaction. Plant Phys. 2006, 142, 1304–1317. [Google Scholar] [CrossRef] [PubMed]
  45. Remans, T.; Nacry, P.; Pervent, M.; Girin, T.; Tillard, P.; Lepetit, M.; Gojon, A. Acentral role for the nitrate transporter NRT21 in the integrated morphological and physiological responses of the root system to nitrogen limitation in Arabidopsis. Plant Phys. 2006, 140, 909–921. [Google Scholar] [CrossRef]
  46. Li, W.B.; Wang, Y.; Okamoto, M.; Crawford, N.M.; Siddiqi, M.Y.; Glass, A.D.M. Dissection of the AtNRT21: AtNRT22 inducible high-affinity nitrate transporter gene cluster. Plant Phys. 2007, 143, 425–433. [Google Scholar] [CrossRef] [PubMed]
  47. Kiba, T.; Feria-Bourrellier, A.B.; Lafouge, F.; Lezhneva, L.; Boutet-Mercey, S.; Orsel, M.; Brehaut, V.; Miller, A.; Daniel-Vedele, F.; Sakakibara, H. The Arabidopsis nitrate transporter NRT24 plays a double role in roots and shoots of nitrogen-starved plants. Plant Cell 2012, 24, 245–258. [Google Scholar] [CrossRef]
  48. Lezhneva, L.; Kiba, T.; Feria-Bourrellier, A.B.; Lafouge, F.; Boutet-Mercey, S.; Zoufan, P.; Sakakibara, H.; Daniel-Vedele, F.; Krapp, A. The Arabidopsis nitrate transporter NRT25 plays a role in nitrate acquisition and remobilization in nitrogen-starved plants. Plant J. 2014, 80, 230–241. [Google Scholar] [CrossRef] [PubMed]
  49. Hill, P.W.; Quilliam, R.S.; DeLuca, T.H.; Farrar, J.; Farrell, M.; Roberts, P.; Newsham, K.K.; Hopkins, D.W.; Bardgett, R.D.; Jones, D.L. Acquisition and assimilation of nitrogen as peptide-bound and D-enantiomers of amino acids by wheat. PLoS ONE 2011, 6, e19220. [Google Scholar] [CrossRef] [PubMed]
  50. Paungfoo-Lonhienne, C.; Lonhienne, T.G.A.; Rentsch, D.; Schmidt, S. Plants can use protein as a nitrogen source without assistance from other organisms. Prod. Natl. Acad. Sci. USA 2008, 105, 4524–4529. [Google Scholar] [CrossRef] [PubMed]
  51. Näsholm, T.; Kielland, K.; Ganeteg, U. Uptake of organic nitrogen by plants. New Phytol. 2009, 182, 31–48. [Google Scholar] [CrossRef]
  52. Lipson, D.; Näsholm, T. The unexpected versatility of plants: Organic nitrogen use and availability in terrestrial ecosystems. Oecologia 2001, 128, 305–316. [Google Scholar] [CrossRef] [PubMed]
  53. Raab, T.K.; Lipson, D.A.; Monson, R.K. Soil amino acid utilization among species of the cyperaceae: Plant and soil processes. Ecology 1999, 80, 2408–2419. [Google Scholar] [CrossRef]
  54. Schulten, H.-R.; Schnitzer, M. The chemistry of soil organic nitrogen: A review. Biol. Fertil. Soils 1998, 26, 1–15. [Google Scholar] [CrossRef]
  55. Rentsch, D.; Schmidt, S.; Tegeder, M. Transporters for uptake and allocation of organic nitrogen compounds in plants. FEBS Lett. 2007, 581, 2281–2289. [Google Scholar] [CrossRef] [PubMed]
  56. Muratore, C.; Espen, L.; Prinsi, B. Nitrogen uptake in plants: The plasma membrane root transport systems from a physiological proteomic perpective. Plants 2021, 10, 681. [Google Scholar] [CrossRef]
  57. Yang, G.; Wei, Q.; Huang, H.; Xia, J. Amino acid transporters in plant cells: A brief review. Plants 2020, 9, 967. [Google Scholar] [CrossRef]
  58. Greenfield, L.M.; Hill, P.W.; Paterson, E.; Baggs, E.M.; Jones, D.L. Do plants use root-derived proteases to promote the uptake of soil organic nitrogen? Plant Soil 2020, 456, 355–367. [Google Scholar] [CrossRef] [PubMed]
  59. Lea, P.J.; Miflin, B.J. Nitrogen assimilation and its relevance to crop improvement. Annu. Plant Rev. Online 2018, 42, 1–40. [Google Scholar]
  60. Phillips, D.A.; Fox, T.C.; Six, J. Root exudation (net efflux of amino acids) may increase rhizodeposition under elevated CO2. Glob. Change Biol. 2006, 12, 561–567. [Google Scholar] [CrossRef]
  61. Contreras-Cornejo, H.A.; Macías-Rodríguez, L.; Del-Val, E.; Larsen, J. Ecological functions of Trichoderma spp. and their secondary metabolites in the rhizosphere: Interactions with plants. FEMS Microbiol Ecol. 2016, 92, fiw036. [Google Scholar] [CrossRef]
  62. Warren, C.R. Organic N molecules in the soil solution: What is known, what is unknown and the path forwards. Soil Biol. Biochem. 2014, 375, 1–19. [Google Scholar] [CrossRef]
  63. Warren, C.R. Variation in small organic N compounds and amino acid enantiomers along an altitudinal gradient. Soil Biol. Biochem. 2017, 115, 197–212. [Google Scholar] [CrossRef]
  64. Nardi, S.; Schiavon, M.; Francioso, O. Chemical Structure and Biological Activity of Humic Substances Define Their Role as Plant Growth Promoters. Molecules 2021, 26, 2256. [Google Scholar] [CrossRef] [PubMed]
  65. Trevisan, S.; Francioso, O.; Quaggiotti, S.; Nardi, S. Humic substances biological activity at the plant-soil interface: From environmental aspects to molecular factors. Plant Signal. Behav. 2010, 5, 635–643. [Google Scholar] [CrossRef]
  66. Roberts, P.; Jones, D. Microbial and plant uptake of free amino sugars in grassland soils. Soil Biol. Biochem. 2012, 49, 139–149. [Google Scholar] [CrossRef]
  67. Glanville, H.; Hill, P.; Schnepf, A.; Oburger, E.; Jones, D. Combined use of empirical data and mathematical modelling to better estimate the microbial turnover of isotopically labelled carbon substrates in soil. Soil Biol.Biochem. 2016, 94, 154–168. [Google Scholar] [CrossRef]
  68. Farzadfar, S.; Knight, J.D.; Congreves, K.A. Soil organic nitrogen: An overlooked but potentially significant contribution to crop nutrition. Plant Soil 2021, 462, 7–23. [Google Scholar] [CrossRef]
  69. Moreau, D.; Pivato, B.; Bru, D.; Busset, H.; Deau, F.; Faivre, C.; Matejicek, A.; Strbik, F.; Philippot, L.; Mougel, C. Plant traits related to nitrogen uptake influence plant-microbe competition. Ecology 2015, 96, 2300–2310. [Google Scholar] [CrossRef] [PubMed]
  70. Berendsen, R.L.; Pieterse, C.M.; Bakker, P.A. The rhizosphere microbiome and plant health. Trends Plant Sci. 2012, 17, 478–486. [Google Scholar] [CrossRef]
  71. Philippot, L.; Raaijmakers, J.M.; Lemanceau, P.; Van Der Putten, W.H. Going back to the roots: The microbial ecology of the rhizosphere. Nat Rev. Microbiol. 2013, 11, 789–799. [Google Scholar] [CrossRef]
  72. de Vries, F.T.; Bardgett, R.D. Plant community controls on short-term ecosystem nitrogen retention. New Phytol. 2016, 210, 861–874. [Google Scholar] [CrossRef] [PubMed]
  73. Abalos, D.; van Groenigen, J.W.; De Deyn, G.B. What plant functional traits can reduce nitrous oxide emissions from intensively managed grasslands? Glob. Chang. Biol. 2018, 24, e248–e258. [Google Scholar] [CrossRef]
  74. Cantarel, A.A.M.; Pommier, T.; Desclos-Theveniau, M.; Diquelou, S.; Dumont, M.; Grassein, F.; Kastl, E.-M.; Grigulis, K.; Laine, P.; Lavorel, S.; et al. Using plant traits to explain plant–microbe relationships involved in nitrogen acquisition. Ecology 2015, 96, 788–799. [Google Scholar] [CrossRef]
  75. Roumet, C.; Birouste, M.; Picon-Cochard, C.; Ghestem, M.; Osman, N.; Vrignon-Brenas, S.; Cao, K.; Stokes, A. Root structure–function relationships in 74 species: Evidence of a root economics spectrum related to carbon economy. New Phytol. 2016, 210, 815–826. [Google Scholar] [CrossRef] [PubMed]
  76. Wright, I.J.; Reich, P.B.; Westoby, M.; Ackerly, D.D.; Baruch, Z.; Bongers, F.; Cavender-Bares, J.; Chapin, T.; Cornelissen, J.H.C.; Diemer, M.; et al. The worldwide leaf economics spectrum. Nature 2004, 428, 821–827. [Google Scholar] [CrossRef]
  77. Harty, M.A.; Forrestal, P.J.; Watson, C.J.; McGeough, K.L.; Carolan, R.; Elliot, C.; Krol, D.; Laughlin, R.J.; Richards, K.G.; Lanigan, G.J. Reducing nitrous oxide emissions by changing N fertiliser use from calcium ammonium nitrate (CAN) to urea based formulations. Sci. Total Environ. 2016, 563, 576–586. [Google Scholar] [CrossRef]
  78. Halvorson, A.D.; Snyder, C.S.; Blaylock, A.D.; Del Grosso, S.J. Enhanced-efficiency nitrogen fertilizers: Potential role in nitrous oxide emission mitigation. J. Agron. 2014, 106, 715–722. [Google Scholar] [CrossRef]
  79. Cameron, K.C.; Di, H.J.; Moir, J.L. Nitrogen losses from the soil/plant system: A review. Ann. Appl. Biol. 2013, 162, 145–173. [Google Scholar] [CrossRef]
  80. Li, Q.; Cui, X.; Liu, X.; Roescke, M.; Pasda, G.; Zerulla, W.; Wissemeier, A.H.; Chen, X.; Goulding, K.; Zhang, F. A new urease-inhibiting formulation decreases ammonia volatilization and improves maize nitrogen utilization in North China Plain. Sci. Rep. 2017, 7, 43853. [Google Scholar] [CrossRef] [PubMed]
  81. Heffer, P.; Prud’homme, M. Fertilizer outlook 2013–2017. In Proceedings of the 81st International Fertilizer Industry Association Conference, Chicago, IL, USA, 13 March 2013; Paper No.: A/13/78. p. 8. [Google Scholar]
  82. Mazzei, L.; Cianci, M.; Contaldo, U.; Ciurli, S. Insights into urease inhibition by N-(n-Butyl) phosphoric triamide through an integrated structural and kinetic approach. J. Agric. Food Chem. 2019, 67, 2127–2138. [Google Scholar] [CrossRef]
  83. Svane, S.; Sigurdarson, J.J.; Finkenwirth, F.; Eitinger, T.; Karring, H. Inhibition of urease activity by different compounds provides insight into the modulation and association of bacterial nickel import and ureolysis. Sci. Rep. 2020, 10, 1–14. [Google Scholar]
  84. Chen, A.; Lei, B.; Hu, W.; Lu, Y.; Mao, Y.; Duan, Z.; Shi, Z. Characteristics of ammonia volatilization on rice grown under different nitrogen application rates and its quantitative predictions in Erhai Lake Watershed, China. Nutr. Cycl. Agroecosys. 2015, 101, 139–152. [Google Scholar] [CrossRef]
  85. Byrne, M.P.; Tobin, J.T.; Forrestal, P.J.; Danaher, M.; Nkwonta, C.G.; Richards, K.; Cummins, E.; Hogan, S.A.; O’Callaghan, T.F. Urease and nitrification inhibitors—As mitigation tools for greenhouse gas emissions in sustainable dairy systems: A review. Sustainability 2020, 12, 6018. [Google Scholar] [CrossRef]
  86. Hayatsu, M.; Tago, K.; Saito, M. Various players in the nitrogen cycle: Diversity and functions of the microorganisms involved in nitrification and denitrification. J. Soil Sci. Plant Nutr. 2008, 54, 33–45. [Google Scholar] [CrossRef]
  87. Prosser, J.I.; Nicol, G.W. Archaeal and bacterial ammonia-oxidisers in soil: The quest for niche specialisation and differentiation. Trends Microbiol. 2012, 20, 523–531. [Google Scholar] [CrossRef] [PubMed]
  88. Kowalchuk, G.A.; Stephen, J.R. Ammonia-oxidizing bacteria: A model for molecular microbial ecology. Annu. Rev Microbiol. 2001, 55, 485–529. [Google Scholar] [CrossRef] [PubMed]
  89. Daims, H.; Lücker, S.; Wagner, M. A new perspective on microbes formerly known as nitrite-oxidizing bacteria. Trends Microbiol. 2016, 24, 699–712. [Google Scholar] [CrossRef] [PubMed]
  90. van Kessel, M.A.; Speth, D.R.; Albertsen, M.; Nielsen, P.H.; Op den Camp, H.J.M.; Kartal, B.; Jetten, M.S.M.; Lucker, S. Complete nitrification by a single microorganism. Nature 2015, 528, 555–559. [Google Scholar] [CrossRef]
  91. Seitzinger, S.; Harrison, J.A.; Bohlke, J.K.; Baouwman, A.F.; Lowrance, R.; Peterson, B.; Tobias, C.; Van Drecht, G. Denitrification across landscapes and waterscapes: A synthesis. Ecol. Appl. 2006, 16, 2064–2090. [Google Scholar] [CrossRef]
  92. Masson-Delmotte, V.; Zhai, P.; Pirani, A.; Connors, S.L.; Péan, C.; Berger, S.; Caud, N.; Chen, Y.; Goldfarb, L.; Gomis, M.I.; et al. (Eds.) IPCC Climate Chang. 2021: The Physical Science Basis. Contribution of Working Group I to the Sixth Assessment Report of the Intergovernmental Panel on Climate Change; Cambridge University Press: Cambridge, UK, 2021; in press. [Google Scholar]
  93. Di, H.J.; Cameron, K.C. Nitrate leaching in temperate agroecosystems: Sources, factors and mitigating strategies. Nutr. Cycl. Agroecosyst. 2002, 64, 237–256. [Google Scholar] [CrossRef]
  94. Netz, B.; Davidson, O.; Bosch, P.; Dave, R.; Meyer, L. Climate Change 2007: Mitigation. In Contribution of Working Group III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change; Summary for Policymakers; Intergovernmental Panel on Climate Change (IPCC): Geneva, Switzerland, 2007. [Google Scholar]
  95. Singh, M.; Yadav, D.S.; Kumar, V. Leaching and transformation of urea in dry and wet soils as affected by irrigation water. Plant Soil 1984, 81, 411–420. [Google Scholar] [CrossRef]
  96. Dawar, K.; Zaman, M.; Rowarth, J.S.; Blennerhassett, J.; Turnbull, M.H. Urea hydrolysis and lateral and vertical movement in the soil: Effects of urease inhibitor and irrigation. Biol. Fertil. Soils 2011, 47, 139–146. [Google Scholar] [CrossRef]
  97. Domínguez, M.J.; Sanmartín, C.; Font, M.; Palop, J.A.; San Francisco, S.; Urrutia, O.; Houdusse, F.; García-Mina, J.M. Design, synthesis, and biological evaluation of phosphoramide derivatives as urease inhibitors. J. Agric. Food Chem. 2008, 56, 3721–3731. [Google Scholar] [CrossRef]
  98. Zhou, X.; Wu, L.H.; Dai, F. Influence of a new phosphoramide urease inhibitor on urea-N transformation in different texture soil. J. Appl. Ecol. 2016, 27, 4003–4012. [Google Scholar]
  99. Zerulla, W.; Barth, T.; Dressel, J.; Erhardt, K.; Horchler von Locquenghien, K.; Pasda, G.; Rädle, M.; Wissemeier, A. 3,4-Dimethylpyrazole phosphate (DMPP)–A new nitrification inhibitor for agriculture and horticulture. Biol. Fertil. Soils 2001, 34, 79–84. [Google Scholar] [CrossRef]
  100. Subbarao, G.; Ishikawa, T.; Ito, O.; Nakahara, K.; Wang, H.; Berry, W. A bioluminescence assay to detect nitrification inhibitors released from plant roots: A case study with Brachiaria Humidicola. Plant Soil 2006, 288, 101–112. [Google Scholar] [CrossRef]
  101. Subbarao, G.V.; Nakahara, K.; Hurtado, M.P.; Ono, H.; Moreta, D.E.; Salcedo, A.F.; Yoshihashi, A.T.; Ishikawa, T.; Ishitani, M.; Ohnishi-Kameyama, M.; et al. Evidence for biological nitrification inhibition in Brachiaria pastures. Prod. Natl. Acad. Sci. USA 2009, 106, 17302–17307. [Google Scholar] [CrossRef] [PubMed]
  102. Zakir, H.; Subbarao, G.V.; Pearse, S.J.; Gopalakrishnan, S.; Ito, O.; Ishikawa, T.; Kawano, N.; Nakahara, K.; Yoshihashi, T.; Ono, H.; et al. Detection, isolation and characterization of a root-exuded compound, methyl 3-(4-hydroxyphenyl) propionate, responsible for biological nitrification inhibition by sorghum (sorghum bicolor). New Phytol. 2008, 5, 442–451. [Google Scholar] [CrossRef] [PubMed]
  103. Zhao, M.; Jones, C.M.; Meijer, J.; Lundquist, P.-O.; Fransson, P.; Carlsson, G.; Hallin, S. Intercropping affects genetic potential for inorganic nitrogen cycling by root-associated microorganisms in Medicago sativa and Dactylis glomerata. Appl. Soil Ecol. 2017, 119, 260–266. [Google Scholar] [CrossRef]
  104. Nardi, P.; Laanbroek, H.J.; Nicol, G.W.; Renella, G.; Cardinale, M.; Pietramellara, G.; Weckwerth, W.; Trinchera, A.; Ghatak, A.; Nannipieri, P. Biological nitrification inhibition in the rhizosphere: Determining interactions and impact on microbially mediated processes and potential applications. FEMS Microbiol. Rev. 2020, 44, 874–908. [Google Scholar] [CrossRef]
  105. Subbarao, G.V.; Arango, J.; Masahiro, K.; Hooper, A.M.; Yoshihashi, T.; Ando, Y.; Nakahara, K.; Deshpande, S.; Ortiz-Monasterio, I.; Ishitani, M.; et al. Genetic mitigation strategies to tackle agricultural GHG emissions: The case for biological nitrification inhibition technology. Plant Sci. 2017, 262, 165–168. [Google Scholar] [CrossRef]
  106. Subbarao, G.V.; Sahrawat, K.L.; Nakahara, K.; Ishikawa, T.; Kishii, M.; Rao, I.M.; Hash, C.T.; Geouge, T.S.; Srinivasa Roa, P.; Nardi, P.; et al. Biological nitrification inhibition: A novel strategy to regulate nitrification in agricultural systems. Adv. Agron. 2012, 114, 249–302. [Google Scholar]
  107. Subbarao, G.V.; Sahrawat, K.L.; Nakahara, K.; Rao, I.M.; Ishitani, M.; Hash, C.T.; Kishii, M.; Bonnett, D.G.; Berry, W.L.; Lata, J.C. A paradigm shift towards low-nitrifying production systems: The role of biological nitrification inhibition (BNI). Ann. Bot. 2013, 112, 297–316. [Google Scholar] [CrossRef]
  108. Wang, X.; Bai, J.; Xie, T.; Wang, W.; Zhang, G.; Yin, S.; Wang, D. Effects of biological nitrification inhibitors on nitrogen use efficiency and greenhouse gas emissions in agricultural soils: A review. Ecotoxicol. Environ. Saf. 2021, 220, 112338. [Google Scholar] [CrossRef]
  109. Prasad, R.; Power, J.F. Nitrification inhibitors for agriculture, health, and the environment. Adv. Agron. 1995, 54, 233–281. [Google Scholar]
  110. Dong, G.; Xiao, D.; Zhang, C.-H. The effect of nitrification inhibitors on nitrogen cycle: A comprehensive review. In Proceedings of the IOP Conference Series: Earth and Environmental Science 2020 International Symposium on water, Ecology and Environment, Beijing, China, 6–8 December 2021; Volume 690, p. 012012. [Google Scholar]
  111. Mazzei, L.; Musiani, F.; Ciurli, S. The structure-based reaction mechanism of urease, a nickel dependent enzyme: Tale of a long debate. J. Biol. Inorg. Chem. 2020, 25, 829–845. [Google Scholar] [CrossRef] [PubMed]
  112. Lasisi, A.A.; Akinremi, O.O. Kinetics and thermodynamics of urea hydrolysis in the presence of urease and nitrification inhibitors. Can. J. Soil Sci. 2021, 101, 192–202. [Google Scholar] [CrossRef]
  113. Suter, H.C.; Pengthamkeerati, P.; Walker, C.; Chen, D. Influence of temperature and soil type on inhibition of urea hydrolysis by N-(n-butyl) thiophosphoric triamide in wheat and pasture soils in south-eastern Australia. Soil Res. 2011, 49, 315–319. [Google Scholar] [CrossRef]
  114. Klimczyk, M.; Siczek, A.; Schimmelpfennig, L. Improving the efficiency of urea-based fertilization leading to reduction in ammonia emission. Sci. Total Environ. 2021, 771, 145483. [Google Scholar] [CrossRef] [PubMed]
  115. Kafarski, P.; Talma, M. Recent advances in design of new urease inhibitors: A review. J. Adv. Res. 2018, 13, 101–112. [Google Scholar] [CrossRef]
  116. Modolo, L.V.; da-Silva, C.J.; Brandão, D.S.; Chaves, I.S. A minireview on what we have learned about urease inhibitors of agricultural interest since mid-2000s. J. Adv. Res. 2018, 13, 29–37. [Google Scholar] [CrossRef]
  117. Brito, T.O.; Souza, A.X.; Mota, Y.C.C.; Morais, V.S.S.; De Souza, L.T.; De Fátima, Â.; Macedo, F.; Modolo, L.V. Design, syntheses and evaluation of benzoylthioureas as urease inhibitors of agricultural interest. RSC Adv. 2015, 5, 44507–44515. [Google Scholar] [CrossRef]
  118. Imran, M.; Waqar, S.; Ogata, K.; Ahmed, M.; Noreen, Z.; Javed, S.; Bibi, N.; Bokhari, H.; Amjad, A.; Muddassar, M. Identification of novel bacterial urease inhibitors through molecular shape and structure based virtual screening approaches. RSC Adv. 2020, 10, 16061–16070. [Google Scholar] [CrossRef]
  119. Hanif, M.; Shoaib, K.; Saleem, M.; Hasan Rama, N.; Zaib, S.; Iqbal, J. Synthesis, urease inhibition, antioxidant, antibacterial, and molecular docking studies of 1,3,4-oxadiazole derivatives. ISRN Pharmacol. 2012, 2012, 928901. [Google Scholar] [CrossRef]
  120. Mazzei, L.; Cianci, M.; Musiani, F.; Ciurli, S. Inactivation of urease by 1, 4-benzoquinone: Chemistry at the protein surface. Dalton Trans. 2016, 45, 5455–5459. [Google Scholar] [CrossRef] [PubMed]
  121. Babu, T.M.C.; Rajesh, S.S.; Bhaskar, B.V.; Devi, S.; Rammohan, A.; Sivaraman, T.; Rajendra, W. Molecular docking, molecular dynamics simulation, biological evaluation and 2D QSAR analysis of flavonoids from Syzygium alternifolium as potent anti-Helicobacter pylori agents. RSC Adv. 2017, 7, 18277–18292. [Google Scholar] [CrossRef]
  122. Christianson, C.B.; Howard, R.G. Use of soil thin-layer chromatography to assess the mobility of the phosphoric triamide urease inhibitors and urea in soil. Soil Biol. Biochem. 1994, 26, 1161–1164. [Google Scholar] [CrossRef]
  123. McCarty, G.W. Modes of action of nitrification inhibitors. Biol. Fertil. Soils 1999, 29, 1–9. [Google Scholar] [CrossRef]
  124. Lees, H. Effect of Copper-Enzyme Poisons on Soil Nitrification. Nature 1946, 158, 97. [Google Scholar] [CrossRef]
  125. Yildirim, S.C.; Walker, R.M.; Roessner, U.; Wille, U. Assessing the efficacy, acute toxicity, and binding modes of the agricultural nitrification inhibitors 3, 4-Dimethyl-1 H-pyrazole (DMP) and Dicyandiamide (DCD) with Nitrosomonas europaea. ACS J. Agric. Sci. 2023, 3, 222–231. [Google Scholar] [CrossRef]
  126. Fisher, O.S.; Kenney, G.E.; Ross, M.O.; Ro, S.Y.; Lemma, B.E.; Batelu, S.; Thomas, P.M.; Sosnowski, V.C.; DeHart, C.J.; Kelleher, N.L.; et al. Characterization of a long overlooked copper protein from methane- and ammonia-oxidizing bacteria. Nat. Commun. 2018, 9, 4276. [Google Scholar] [CrossRef]
  127. Lieberman, R.L.; Rosenzweig, A.C. Crystal structure of a membrane-bound metalloenzyme that catalyses the biological oxidation of methane. Nature 2005, 434, 177–182. [Google Scholar] [CrossRef]
  128. Suzuki, I.; Dular, U.; Kwok, S.C. Ammonia or ammonium ion as substrate for oxidation by Nitrosomonas europaea cells and extracts. J. Bacteriol. 1974, 120, 556–558. [Google Scholar] [CrossRef]
  129. O’Sullivan, C.A.; Duncan, E.G.; Whisson, K.; Treble, K.; Ward, P.R.; Roper, M.M. A colourimetric microplate assay for simple, high throughput assessment of synthetic and biological nitrification inhibitors. Plant Soil 2017, 413, 275–287. [Google Scholar] [CrossRef]
  130. Leininger, S.; Urich, T.; Schloter, M.; Schwark, L.; Qi, J.; Nicol, G.W.; Prosser, J.I.; Schuster, S.; Schleper, C. Archaea predominate among ammonia-oxidizing prokaryotes in soils. Nature 2006, 442, 806–809. [Google Scholar] [CrossRef] [PubMed]
  131. Wright, C.L.; Schatteman, A.; Crombie, A.T.; Murrell, J.C.; Lehtovirta-Morley, L.E. Inhibition of ammonia monooxygenase from ammonia-oxidizing archaea by linear and aromatic alkynes. Appl. Environ. Microbiol. 2020, 86, e02388-19. [Google Scholar] [CrossRef]
  132. Kaur-Bhambra, J.; Wardak, D.L.R.; Prosser, J.I.; Gubry-Rangin, C. Revisiting plant biological nitrification inhibition efficiency using multiple archaeal and bacterial ammonia-oxidising cultures. Biol Fertil. Soils 2021, 58, 241–249. [Google Scholar] [CrossRef]
  133. Bédard, C.; Knowles, R. Physiology, biochemistry, and specific inhibitors of CH4, NH4+, and CO oxidation by methanotrophs and nitrifiers. Microbiol. Rev. 1989, 53, 68–84. [Google Scholar] [CrossRef] [PubMed]
  134. Juliette, L.Y.; Hyman, M.R.; Arp, D.J. Inhibition of ammonia oxidation in Nitrosomonas europaea by sulfur compounds: Thioethers are oxidized to sulfoxides by ammonia monooxygenase. Appl. Environ. Microbiol. 1993, 59, 3718–3727. [Google Scholar] [CrossRef]
  135. Hubley, J.H.; Mitton, J.R.; Wilkinson, J.F. The oxidation of carbon monoxide by methane-oxidizing bacteria. Arch. Microbiol. 1974, 95, 365–368. [Google Scholar] [CrossRef]
  136. Hyman, M.R.; Wood, P.M. Methane oxidation by Nitrosomonas europaea. J. Biochem. 1983, 212, 31–37. [Google Scholar] [CrossRef] [PubMed]
  137. Hyman, M.R.; Wood, P.M. Ethylene oxidation by Nitrosomonas europaea. Arch. Microbiol. 1984, 137, 155–158. [Google Scholar] [CrossRef]
  138. Stirling, D.I.; Dalton, H. The fortuitous oxidation and cometabolism of various carbon compounds by whole-cell suspensions of Methylococcus capsulatus(Bath). FEMS Microbiol Lett. 1979, 5, 315–318. [Google Scholar] [CrossRef]
  139. Voysey, P.A.; Wood, P.M. Methanol and formaldehyde oxidation by an autotrophic nitrifying bacterium. Microbiol. 1987, 133, 283–290. [Google Scholar] [CrossRef]
  140. Ward, B.B. Kinetic studies on ammonia and methane oxidation by Nitrosococcus oceanus. Arch. Microbiol. 1987, 147, 126–133. [Google Scholar] [CrossRef]
  141. Hyman, M.R.; Sansome-Smith, A.W.; Shears, J.H.; Wood, P.M. A kinetic study of benzene oxidation to phenol by whole cells of Nitrosomonas europaea and evidence for the further oxidation of phenol to hydroquinone. Arch. Microbiol. 1985, 143, 302–306. [Google Scholar] [CrossRef]
  142. Rasche, M.E.; Hicks, R.E.; Hyman, M.R.; Arp, D.J. Oxidation of monohalogenated ethanes and n-chlorinated alkanes by whole cells of Nitrosomonas europaea. J. Bacteriol. 1990, 172, 5368–5373. [Google Scholar] [CrossRef]
  143. Fernández, M.L.; Estrin, D.A.; Bari, S.E. Theoretical insight into the hydroxylamine oxidoreductase mechanism. J. Inorg. Biochem. 2008, 102, 1523–1530. [Google Scholar] [CrossRef] [PubMed]
  144. Whittaker, M.; Bergmann, D.; Arciero, D.; Hooper, A.B. Electron transfer during the oxidation of ammonia by the chemolithotrophic bacterium Nitrosomonas europaea. Biochem. Biophys. Acta. 2000, 1459, 346–355. [Google Scholar] [CrossRef] [PubMed]
  145. Lawton, T.J.; Ham, J.; Sun, T.; Rosenzweig, A.C. Structural conservation of the B subunit in the ammonia monooxygenase/particulate methane monooxygenase superfamily. Proteins 2014, 82, 2263–2267. [Google Scholar] [CrossRef]
  146. Mario, C.M.; Carmen, G.M.; Adrián, B.L.; Luis, L.; Beñat, A. Mechanism of action of nitrification inhibitors based on dimethylpyrazole: A matter of chelation. Sci. Total Environ. 2021, 752, 141885. [Google Scholar]
  147. Ensign, S.A.; Hyman, M.R.; Arp, D.J. In vitro activation of ammonia monooxygenase from Nitrosomonas europaea by copper. J. Bacteriol. 1993, 175, 1971–1980. [Google Scholar] [CrossRef]
  148. Huérfano, X.; Fuertes-Mendizábal, T.; Fernández-Diez, K.; Estavillo, J.M.; González-Murua, C.; Menéndez, S. The new nitrification inhibitor 3, 4-dimethylpyrazole succinic (DMPSA) as an alternative to DMPP for reducing N2O emissions from wheat crops under humid Mediterranean conditions. Eur. J. Agron. 2016, 80, 78–87. [Google Scholar] [CrossRef]
  149. Recio, J.; Alvarez, J.M.; Rodriguez-Quijano, M.; Vallejo, A. Nitrification inhibitor DMPSA mitigated N2O emission and promoted NO sink in rainfed wheat. Environ. Pollut. 2019, 245, 199–207. [Google Scholar] [CrossRef]
  150. Gao, H.; Li, J.; Xu, F. Synthesis of a novel polymer nitrification inhibitor with acrylic acid and 3,4-Dimethylpyrazole. J. Agric. Food Chem. 2021, 69, 307–3311. [Google Scholar] [CrossRef] [PubMed]
  151. Vejan, P.; Khadiran, T.; Abdullah, R.; Ahmad, N. Controlled release fertilizer: A review on developments, applications and potential in agriculture. J. Control. Release 2021, 339, 321–334. [Google Scholar] [CrossRef]
  152. Asgedom, H.; Tenuta, M.; Flaten, D.N.; Gao, X.; Kebreab, E. Nitrous oxide emissions from a clay soil receiving granular urea formulations and dairy manure. Agron. J. 2014, 106, 732–774. [Google Scholar] [CrossRef]
  153. Dougherty, W.; Collins, D.; Zwieten, L.V.; Rowlings, D. Nitrification (DMPP) and urease (NBPT) inhibitors had no effect on pasture yield, nitrous oxide emissions, or nitrate leaching under irrigation in a hot-dry climate. Soil Res. 2016, 54, 675–683. [Google Scholar] [CrossRef]
  154. Guardia, G.; Marsdden, K.A.; Vallejo, A.; Jones, D.L.; Chadwick, D.R. Determining the influence of environmental and edaphic factors on the fate of the nitrification inhibitors DCD and DMPP in soil. Sci. Total. Environ. 2018, 624, 1202–1212. [Google Scholar] [CrossRef] [PubMed]
  155. Lan, T.; Suter, H.; Liu, R.; Yuan, S.; Chen, D. Effects of nitrification inhibitors on gross N nitrification rate, ammonia oxidizers, and N2O production under different temperatures in two pasture soils. Environ. Sci. Pollut. Res. 2018, 25, 28344–28354. [Google Scholar] [CrossRef]
  156. Sha, Z.; Li, Q.; Lv, T.; Misselbrook, T.; Liu, X. Response of ammonia volatilization to biochar addition: A meta-analysis. Sci. Total Environ. 2019, 655, 1387–1396. [Google Scholar] [CrossRef]
  157. Sha, Z.; Ma, X.; Wang, J.; Lv, T.; Li, Q.; Misselbrook, T.; Liu, X. Effect of N stabilizers on fertilizer-N fate in the soil-crop system: A meta-analysis. Agric. Ecosyst. Environ. 2020, 290, 106763. [Google Scholar] [CrossRef]
  158. Zhu, G.; Ju, X.; Zhang, J.; Müller, C.; Sylvester-Bradley, R. Effects of the nitrification inhibitor DMPP (3,4-dimethylpyrazole phosphate) on gross N transformation rates and N2O emissions. Biol. Fertil. Soils 2019, 55, 603–615. [Google Scholar] [CrossRef]
  159. Carmona, G.; Christianson, C.B.; Byrnes, B.H. Temperature and low concentration effects of the urease inhibitor N-(n-butyl) thiophosphoric triamide (n-BTPT) on ammonia volatilization from urea. Soil Biol. Biochem. 1990, 22, 933–937. [Google Scholar] [CrossRef]
  160. Moyo, C.; Kissel, D.; Cabrera, M. Temperature effects on soil urease activity. Soil Biol. Biochem. 1989, 21, 935–993. [Google Scholar] [CrossRef]
  161. McGeough, K.L.; Watson, C.J.; Müller, C.; Laughlin, R.J.; Chadwick, D.R. Evidence that the efficacy of the nitrification inhibitor dicyandiamide (DCD) is affected by soil properties in UK soils. Soil Biol. Biochem. 2016, 94, 222–232. [Google Scholar] [CrossRef]
  162. Kelliher, F.; van Koten, C.; Kear, M.; Sprosen, M.; Ledgard, S.; de Klein, C.; Letica, S.; Luo, J.; Rys, G. Effect of temperature on dicyandiamide (DCD) longevity in pastoral soils under field conditions. Agric. Ecosyst. Environ. 2014, 186, 201–204. [Google Scholar] [CrossRef]
  163. Irigoyen, I.; Muro, J.; Azpilikueta, M.; Aparicio-Tejo, P.; Lamsfus, C. Ammonium oxidation kinetics in the presence of nitrification inhibitors DCD and DMPP at various temperatures. Aust. J. Soil Res. 2003, 41, 1177–1183. [Google Scholar] [CrossRef]
  164. Mahmood, T.; Rehmat, A.; Asma, L.; Muhammad, S. 4-Amino-1, 2, 4-triazole can be more effective than commercial nitrification inhibitors at high soil temperatures. Soil Res. 2017, 55, 715–722. [Google Scholar] [CrossRef]
  165. Marsden, K.A.; Jones, D.L.; Chadwick, D.R. DMPP is ineffective at mitigating N2O emissions from sheep urine patches in a UK grassland under summer conditions. Agric. Ecosyst. Environ. 2017, 246, 1–11. [Google Scholar] [CrossRef]
  166. Engel, R.; Williams, E.; Wallander, R.; Hilmer, J. Apparent persistence of N-(n-butyl) thiophosphoric triamide is greater in alkaline soils. Soil Sci. Soc. Am. J. 2013, 77, 1424–1429. [Google Scholar] [CrossRef]
  167. Engel, R.E.; Towey, B.D.; Gravens, E. Degradation of the urease inhibitor NBPT as affected by soil pH. Soil Sci. Soc. Am. J. 2015, 79, 1674–1683. [Google Scholar] [CrossRef]
  168. Tao, L.; Yuanliang, S.; Xuewen, L.; Guolin, L. Degradation and its affecting factors of NBPT in soil. Chin. J. Ecol. 2006, 25, 1082–1086. [Google Scholar]
  169. He, J.-Z.; Hu, H.-W.; Zhang, L.-M. Current insights into the autotrophic thaumarchaeal ammonia oxidation in acidic soils. Soil Biol. Biochem. 2012, 55, 146–154. [Google Scholar] [CrossRef]
  170. He, J.-Z.; Shen, J.-P.; Zhang, L.-M.; Di, H.J. A review of ammonia-oxidizing bacteria and archaea in Chinese soils. Front. Microbiol. 2012, 3, 296. [Google Scholar]
  171. Jia, Z.; Zhou, X.; Xia, W.; Fornara, D.; Wang, B.; Wasson, E.A.; Christie, P.; Polz, M.F.; Myrold, D.D. Evidence for niche differentiation of nitrifying communities in grassland soils after 44 years of different field fertilization scenarios. Pedosphere 2020, 30, 87–97. [Google Scholar] [CrossRef]
  172. Liu, R.; Hayden, H.L.; Hu, H.; He, J.; Suter, H.; Chen, D. Effects of the nitrification inhibitor acetylene on nitrous oxide emissions and ammonia-oxidizing microorganisms of different agricultural soils under laboratory incubation conditions. Appl. Soil Ecol. 2017, 119, 80–90. [Google Scholar] [CrossRef]
  173. Prosser, J.I.; Hink, L.; Gubry-Rangin, C.; Nicol, G.W. Nitrous Oxide production by ammonia oxidizers: Physiological diversity, niche differentiation and potential mitigation strategies. Glob. Chang. Biol. 2020, 26, 103–118. [Google Scholar] [CrossRef]
  174. Chen, Q.; Qi, L.; Bi, Q.; Dai, P.; Sun, D.; Sun, C.; Liu, W.; Lu, L.; Ni, W.; Lin, X. Comparative effects of 3,4-dimethylpyrazole phosphate (DMPP) and dicyandiamide (DCD) on ammonia-oxidizing bacteria and archaea in a vegetable soil. Appl. Microbiol. Biotechnol. 2015, 99, 477–487. [Google Scholar] [CrossRef]
  175. Martens-Habbena, W.; Berube, P.M.; Urakawa, H.; De La Torre, J.R.; Stahl, D.A. Ammonia oxidation kinetics determine niche separation of nitrifying Archaea and Bacteria. Nature 2009, 461, 976–979. [Google Scholar] [CrossRef]
  176. Zhang, L.-M.; Hu, H.-W.; Shen, J.-P.; He, J.-Z. Ammonia-Oxidizing archaea have more important role than ammonia-oxidizing bacteria in ammonia oxidation of strongly acidic soils. ISME J. 2012, 6, 1032–1045. [Google Scholar] [CrossRef]
  177. Shi, X.; Hu, H.; He, J.; Chen, D.; Suter, H. Effects of 3, 4-dimethylpyrazole phosphate (DMPP) on nitrification and the abundance and community composition of soil ammonia oxidizers in three land uses. Biol. Fertil. Soils 2016, 52, 927–939. [Google Scholar] [CrossRef]
  178. Oertel, C.; Matschullat, J.; Zurba, K.; Zimmermann, F.; Erasmi, S. Greenhouse gas emissions from soils A review. Chem. Erde-Geochem. 2016, 76, 327–352. [Google Scholar] [CrossRef]
  179. Christianson, C.B.; Baethgen, W.E.; Carmona, G.; Howard, R.G. Microsite reactions of urea-nBTPT fertilizer on the soil surface. Soil Biol. Biochem. 1993, 25, 1107–1117. [Google Scholar] [CrossRef]
  180. Zhang, H.J.; Wu, Z.J.; Zhou, Q.X. Dicyandiamide sorption-desorption behaviour on soils and peat humus. Pedosphere 2004, 14, 395–399. [Google Scholar]
  181. Wu, S.; Wu, L.; Shi, Q.; Wang, Z.; Chen, X.; Li, Y. Effects of a new nitrification inhibitor 3, 4–dimethylpyrazole phosphate (DMPP) on nitrate and potassium leaching in two soils. J. Environ. Sci. 2007, 19, 841–847. [Google Scholar] [CrossRef]
  182. Barth, G.; Von Tucher, S.; Schmidhalter, U. Effectiveness of 3, 4-dimethylpyrazole phosphate as nitrification inhibitor in soil as influenced by inhibitor content, application form, and soil matric potential. Pedosphere 2008, 18, 378–385. [Google Scholar] [CrossRef]
  183. Chen, D.; Suter, H.C.; Islam, A.; Edis, R. Influence of nitrification inhibitors on nitrification and nitrous oxide (N2O) emission from a clay loam soil fertilized with urea. Soil Biol. Biochem. 2010, 42, 660–664. [Google Scholar] [CrossRef]
  184. International Fertilizer Industry Association. Sustainable Management of the Nitrogen Cycle in Agriculture and Mitigation of Reactive Nitrogen Side Effects; IFIA: Paris, France, 2007. [Google Scholar]
  185. Suter, H.C.; Sultana, H.; Davies, R.; Walker, C.; Chen, D. Influence of enhanced efficiency fertilisation techniques on nitrous oxide emissions and productivity response from urea in a temperate Australian ryegrass pasture. Soil Res. 2016, 54, 523–532. [Google Scholar] [CrossRef]
  186. Fan, X.; Yin, C.; Yan, G.; Cui, P.; Shen, Q.; Wang, Q.; Chen, H.; Zhang, N.; Ye, M.; Zhao, Y.; et al. The contrasting effects of N-(n-butyl) thiophosphoric triamide (NBPT) on N2O emissions in arable soils differing in pH are underlain by complex microbial mechanisms. Sci. Total Environ. 2018, 642, 155–167. [Google Scholar] [CrossRef]
  187. Lam, S.K.; Suter, H.; Bai, M.; Walker, C.; Davies, R.; Mosier, A.R.; Chen, D. Using urease and nitrification inhibitors to decrease ammonia and nitrous oxide emissions and improve productivity in a subtropical pasture. Sci. Total Environ. 2018, 644, 1531–1535. [Google Scholar] [CrossRef]
  188. Castellano-Hinojosa, A.; González-López, J.; Vallejo, A.; Bedmar, E.J. Effect of urease and nitrification inhibitors on ammonia volatilization and abundance of N-cycling genes in an agricultural soil. J. Plant Nutr. Soil Sci. USA 2019, 183, 99–109. [Google Scholar] [CrossRef]
  189. Frame, W. Ammonia volatilization from urea treated with NBPT and two nitrification inhibitors. Agron. J. 2017, 109, 378–387. [Google Scholar] [CrossRef]
  190. Pan, B.; Lam, S.K.; Mosier, A.; Luo, Y.; Chen, D. Ammonia volatilization from synthetic fertilizers and its mitigation strategies: A global synthesis. Agric. Ecosyst. Environ. 2016, 232, 283–289. [Google Scholar] [CrossRef]
  191. Soares, J.R.; Cantarella, H.; de Campos Menegale, M.L. Ammonia volatilization losses from surface-applied urea with urease and nitrification inhibitors. Soil Biol. Biochem. 2012, 52, 82–89. [Google Scholar] [CrossRef]
  192. Cantarella, H.; Otto, R.; Soares, J.R.; Silva, A.G.; de Brito, S. Agronomic efficiency of NBPT as a urease inhibitor: A review. J. Adv. Res. 2018, 13, 19–27. [Google Scholar] [CrossRef]
  193. Cui, L.; Li, D.; Wu, Z.; Xue, Y.; Xiao, F.; Zhang, L.; Song, Y.; Li, Y.; Zheng, Y.; Zhang, J.; et al. Effects of nitrification inhibitors on soil nitrification and ammonia volatilization in three soils with different pH. Agronomy 2021, 11, 1674. [Google Scholar] [CrossRef]
  194. Drury, C.F.; Yang, X.; Reynolds, W.D.; Calder, W.; Oloya, T.O.; Woodley, A. combining urease and nitrification inhibitors with incorporation reduces ammonia and nitrous oxide emissions and increases corn yields. J. Environ. Qual. 2017, 46, 939–949. [Google Scholar] [CrossRef] [PubMed]
  195. Grant, C.A.; Bailey, L.D. Effect of seed-placed urea fertilizer and N-(n-butyl) thiophosphoric triamide (NBPT) on emergence and grain yield of barley. Can. J. Plant Sci. 1999, 79, 491–496. [Google Scholar] [CrossRef]
  196. Grant, C.A.; Derksen, D.A.; McLaren, D.; Irvine, R.B. Nitrogen fertilizer and urease inhibitor effects on canola seed quality in a one-pass seeding and fertilizing system. Field Crop. Res. 2011, 121, 201–208. [Google Scholar] [CrossRef]
  197. Xiaobin, W.; Jingfeng, X.; Grant, C.A.; Bailey, L.D. Effects of placement of urea with a urease inhibitor on seedling emergence, N uptake and dry matter yield of wheat. Can. J. Plant Sci. 1995, 75, 449–452. [Google Scholar] [CrossRef]
  198. Zanin, L.; Tomasi, N.; Zamboni, A.; Varanini, Z.; Pinton, R. The urease inhibitor NBPT negatively affects DUR3-mediated uptake and assimilation of urea in maize roots. Front. Plant Sci. 2015, 6, 1007. [Google Scholar] [CrossRef] [PubMed]
  199. Artola, E.; Cruchaga, S.; Ariz, I.; Moran, J.F.; Garnica, M.; Houdusse, F.; Garcia-Mina, J.M.; Ignacio, I.; Lasa, B.; Aparicio-Tejo, P.M. Effect of N-(n-butyl) thiophosphorictriamide on urea metabolism and the assimilation of ammonium by Triticum aestivum L. Plant Growth Regul. 2011, 63, 73–79. [Google Scholar] [CrossRef]
  200. Cruchaga, S.; Lasa, B.; Jauregui, I.; González-Murua, C.; Aparicio-Tejo, P.M.; Ariz, I. Inhibition of endogenous urease activity by NBPT application reveals differential N metabolism responses to ammonium or nitrate nutrition in pea plants: A physiological study. Plant Soil 2013, 373, 813–827. [Google Scholar] [CrossRef]
  201. Cruchaga, S.; Artola, E.; Lasa, B.; Ariz, I.; Irigoyen, I.; Moran, J.F.; Aparicio-Tejo, P.M. Short term physiological implications of NBPT application on the N metabolism of Pisum sativum and Spinacea oleracea. J. Plant Physiol. 2011, 168, 329–336. [Google Scholar] [CrossRef]
  202. Krogmeier, M.J.; McCarty, G.W.; Bremner, J.M. Potential phytotoxicity associated with the use of soil urease inhibitors. Proc. Natl. Acad. Sci. USA 1989, 86, 1110–1112. [Google Scholar] [CrossRef]
  203. Watson, C.J.; Miller, H. Short-term effects of urea amended with the urease inhibitor N-(n-butyl) thiophosphoric triamide on perennial ryegrass. Plant Soil 1996, 184, 33–45. [Google Scholar] [CrossRef]
  204. Trenkel, M.E. Slow-and Controlled-Release and Stabilized Fertilizers: An Option for Enhancing Nutrient Use Efficiency in Agriculture; IFA, International Fertilizer Industry Association: Berlin, Germany, 2010. [Google Scholar]
  205. Rodrigues, J.M.; Lasa, B.; Aparicio-Tejo, P.M.; González-Murua, C.; Marino, D. 3,4-Dimethylpyrazole phosphate and 2-(N-3,4-dimethyl-1H-pyrazol-1-yl) succinic acid isomeric mixture nitrification inhibitors: Quantification in plant tissues and toxicity assays. Sci. Total Environ. 2018, 624, 1180–1186. [Google Scholar] [CrossRef]
  206. Kallio, H.; Linko, R.R.; Tikanmäki, E.; Puntari, I. Effect of nitrapyrin on nitrapyrin residues and nitrate content in red beet roots fertilised with urea. J. Sci. Food Agric. 1980, 31, 701–708. [Google Scholar] [CrossRef]
  207. Dawar, K.; Rahman, U.; Alam, S.S.; Tariq, M.; Khan, A.; Fahad, S.; Datta, R.; Danish, S.; Saud, S.; Noor, M. Nitrification inhibitor and plant growth regulators improve wheat yield and nitrogen use efficiency. J. Plant Growth Regul. 2022, 41, 216–226. [Google Scholar] [CrossRef]
  208. Dawar, K.; Sardar, K.; Zaman, M.; Mueller, C.; Alberto, S.C.; Aamir, K.; Borzouei, A.; Perez-Castillo, A.G. Effects of the nitrification inhibitor nitrapyrin and the plant growth regulator gibberellic acid on yield-scale nitrous oxide emission in maize fields under hot climatic conditions. Pedosphere 2021, 31, 323–331. [Google Scholar] [CrossRef]
  209. Linquist, B.A.; Liu, L.J.; van Kessel, C.; van Groenigen, K.J. Enhanced efficiency nitrogen fertilizers for rice systems: Meta-analysis of yield and nitrogen uptake. Field Crops Res. 2013, 154, 246–254. [Google Scholar] [CrossRef]
  210. Abalos, D.; Jeffery, S.; Sanz-Cobena, A.; Guardia, G.; Vallejo, A. Meta-analysis of the effect of urease and nitrification inhibitors on crop productivity and nitrogen use efficiency. Agric. Ecosyst. Environ. 2014, 189, 136–144. [Google Scholar] [CrossRef]
  211. Yang, M.; Fang, Y.; Sun, D.; Shi, Y. Efficiency of two nitrification inhibitors (dicyandiamide and 3, 4-dimethypyrazole phosphate) on soil nitrogen transformations and plant productivity: A meta-analysis. Sci. Rep. 2016, 6, 22075. [Google Scholar] [CrossRef]
  212. Dimkpa, C.O.; Fugice, J.; Singh, U.; Lewis, T.D. Development of fertilizers for enhanced nitrogen use efficiency—Trends and perspectives. Sci. Total Environ. 2020, 731, 139113. [Google Scholar] [CrossRef]
  213. Recio, J.; Montoya, M.; Álvarez, J.M.; Vallejo, A. Inhibitor-coated enhanced-efficiency N fertilizers for mitigating NOX and N2O emissions in a high-temperature irrigated agroecosystem. Agric. For. Meteorol. 2020, 292, 108110. [Google Scholar] [CrossRef]
  214. Werneck, C.G.; Haim, P.G.; Breda, F.A.D.F.; Monte, M.B.D.M.; Bernardi, A.C.D.C.; Mazur, N.; Polidoro, J.C. Ammonia volatilization and agronomical efficiency of a mixture of urea with natural zeolite for rose fertilization. Pesqui. Agropecu. Bras. 2022, 56, 01449. [Google Scholar] [CrossRef]
  215. Meng, Y.; Wang, J.J.; Wei, Z.; Dodla, S.K.; Fultz, L.M.; Gaston, L.A.; Xiao, R.; Park, J.-h.; Scaglia, G. Nitrification inhibitors reduce nitrogen losses and improve soil health in a subtropical pastureland. Geoderma 2021, 388, 114947. [Google Scholar] [CrossRef]
  216. Weiske, A.; Benckiser, G.; Herbert, T.; Ottow, J. Influence of the nitrification inhibitor 3,4-dimethylpyrazole phosphate (DMPP) in comparison to dicyandiamide (DCD) on nitrous oxide emissions, carbon dioxide fluxes and methane oxidation during 3 years of repeated application in field experiments. Biol. Fertil. Soils 2001, 34, 109–117. [Google Scholar]
  217. Jones, J.W.; Antle, J.M.; Basso, B.; Boote, K.J.; Conant, R.T.; Foster, I.; Godfray, H.C.J.; Herrero, M.; Howitt, R.E.; Janssen, S.; et al. Brief history of agricultural systems modeling. Agric. Syst. 2017, 155, 240–254. [Google Scholar] [CrossRef]
  218. Else, H. Need to Make a Molecule? Ask This AI for Instructions. 2018. Available online: https://www.nature.com/articles/d41586-018-03977-w (accessed on 4 February 2023).
  219. Miljković, F.; Rodríguez-Pérez, R.; Bajorath, J. Impact of artificial intelligence on compound discovery, design, and synthesis. ACS Omega 2021, 6, 33293–33299. [Google Scholar] [CrossRef]
  220. Segler, M.H.S.; Preuss, M.; Waller, M.P. Planning chemical syntheses with deep neural networks and symbolic AI. Nature 2018, 555, 604–610. [Google Scholar] [CrossRef]
  221. Musiani, F.; Arnofi, E.; Casadio, R.; Ciurli, S. Structure-based computational study of the catalytic and inhibition mechanisms of urease. J. Biol. Inorg. Chem. 2001, 6, 300–314. [Google Scholar] [CrossRef] [PubMed]
  222. Weatherburn, M.W. Phenol-hypochlorite reaction for determination of ammonia. Anal. Chem 1967, 39, 971–974. [Google Scholar] [CrossRef]
  223. Hyman, M.R.; Arp, D.J. 14C2H2- and 14CO2-labeling studies of the de novo synthesis of polypeptides by Nitrosomonas europaea during recovery from acetylene and light inactivation of ammonia monooxygenase. J. Biol. Chem. 1992, 267, 1534–1545. [Google Scholar] [CrossRef] [PubMed]
  224. Slangen, J.; Kerkhoff, P. Nitrification inhibitors in agriculture and horticulture: A literature review. J. Fert. Res. 1984, 5, 1–76. [Google Scholar] [CrossRef]
  225. Suwanchaikasem, P.; Idnurm, A.; Selby-Pham, J.; Walker, R.; Boughton, B.A. Root-TRAPR: A modular plant growth device to visualize root development and monitor growth parameters, as applied to an elicitor response of Cannabis sativa. Plant Methods 2022, 18, 46. [Google Scholar] [CrossRef]
  226. Sasse, J.; Kant, J.; Cole, B.J.; Klein, A.P.; Arsova, B.; Schlaepfer, P.; Gao, J.; Lewald, K.; Zhalnina, K.; Kosina, S.; et al. Multilab EcoFAB study shows highly reproducible physiology and depletion of soil metabolites by a model grass. New Phytol. 2019, 222, 1149–1160. [Google Scholar] [CrossRef] [PubMed]
  227. Curl, E.; Truelove, B. Root exudates. In The Rhizosphere; Springer: Berlin/Heidelberg, Germany, 1986; pp. 55–92. [Google Scholar]
  228. Marschner, H. Mechanisms of manganese acquisition by roots from soils. In Manganese in Soils and Plants: Proceedings of the International Symposium on ‘Manganese in Soils and Plants’ held at the Waite Agricultural Research Institute; Graham, R.D., Hannam, R.J., Uren, N.C., Eds.; The University of Adelaide, Glen Osmond, South Australia; August 22–26; as an Australian Bicentennial Even; Springer: Dordrecht, Netherlands, 1988; pp. 191–204. [Google Scholar]
  229. Halvorson, A.D.; Del Grosso, S.J.; Reule, C.A. Nitrogen, tillage, and crop rotation effects on nitrous oxide emissions from irrigated cropping systems. J. Environ. Qual. 2008, 37, 1337–1344. [Google Scholar] [CrossRef] [PubMed]
  230. Rozas, H.R.; Echeverría, H.E.; Picone, L.I. Denitrification in maize under no–tillage: Effect of nitrogen rate and application time. Soil Sci. Soc. Am. J. 2001, 65, 1314–1323. [Google Scholar] [CrossRef]
  231. Burton, D.L.; Zebarth, B.J.; Gillam, K.M.; Macleod, J.A. Effect of split application of fertilizer nitrogen on N2O emissions from potatoes. Can. J. Soil Sci. 2008, 88, 229–239. [Google Scholar] [CrossRef]
  232. Zhang, Z.; Yu, Z.; Zhang, Y.; Shi, Y. Split nitrogen fertilizer application improved grain yield in winter wheat (Triticum aestivum L.) via modulating antioxidant capacity and 13C photosynthate mobilization under water-saving irrigation conditions. Ecol. Process. 2021, 10, 1–13. [Google Scholar] [CrossRef]
  233. Guardia, G.; Gonzalez-Murua, C.; Fuertes-Mendizabal, T.; Vallejo, A. The scarcity and distribution of rainfall drove the performance (i.e., mitigation of N oxide emissions, crop yield and quality) of calcium ammonium nitrate management in a wheat crop under rainfed semiarid conditions. Arch. Agron. Soil Sci. 2020, 66, 1827–1844. [Google Scholar] [CrossRef]
Figure 1. Factors affecting the nitrogen (N) forms uptake by plants.
Figure 1. Factors affecting the nitrogen (N) forms uptake by plants.
Agronomy 13 01603 g001
Scheme 1. Hydrolysis of urea.
Scheme 1. Hydrolysis of urea.
Agronomy 13 01603 sch001
Figure 2. Primary pathways for nitrogen losses from the soil/plant system. Urea is hydrolyzed in the soil to NH4+ (left), catalyzed by the enzyme urease. This results in acidic micro-sites with high concentrations of NH4+ which drive the formation of NH3 (given in red square). NH3 is lost to the environment through volatilization. Nitrification involves the oxidation of NH3 to NO3 (left light green background), which can be leached into the groundwater or converted into various gaseous N forms by the process of denitrification (right dark green background). This results in the loss of N from the soil–plant system. Examples of nitrification and urease inhibitors are given in two white boxes (urease on the left side and nitrification on the right side) that help in the slowdown of the catalytic hydrolysis of urea to NH4+ and delaying the oxidation of NH4+ to NO3, respectively. Figure adapted from Byrne et al. [85]. Although archaeal and fungal nitrifiers are known microbial players [86,87], nitrification has long been considered as the domain of two groups of chemoautotrophic bacteria: the ammonia (for example, Nitrosomonas spp. or Nitrosococcus spp.) and nitrite-oxidizing bacteria (for example, Nitrobacter spp.). The process involves three steps (Equation (1)).
Figure 2. Primary pathways for nitrogen losses from the soil/plant system. Urea is hydrolyzed in the soil to NH4+ (left), catalyzed by the enzyme urease. This results in acidic micro-sites with high concentrations of NH4+ which drive the formation of NH3 (given in red square). NH3 is lost to the environment through volatilization. Nitrification involves the oxidation of NH3 to NO3 (left light green background), which can be leached into the groundwater or converted into various gaseous N forms by the process of denitrification (right dark green background). This results in the loss of N from the soil–plant system. Examples of nitrification and urease inhibitors are given in two white boxes (urease on the left side and nitrification on the right side) that help in the slowdown of the catalytic hydrolysis of urea to NH4+ and delaying the oxidation of NH4+ to NO3, respectively. Figure adapted from Byrne et al. [85]. Although archaeal and fungal nitrifiers are known microbial players [86,87], nitrification has long been considered as the domain of two groups of chemoautotrophic bacteria: the ammonia (for example, Nitrosomonas spp. or Nitrosococcus spp.) and nitrite-oxidizing bacteria (for example, Nitrobacter spp.). The process involves three steps (Equation (1)).
Agronomy 13 01603 g002
Figure 3. Hydrolysis of N-(n-butyl) thiophosphoric triamide (NBPT) to N-(n-butyl) phosphoric triamide (NBPTO) in soils (left). Structure of Duromide (ANVOL, Koch) (right).
Figure 3. Hydrolysis of N-(n-butyl) thiophosphoric triamide (NBPT) to N-(n-butyl) phosphoric triamide (NBPTO) in soils (left). Structure of Duromide (ANVOL, Koch) (right).
Agronomy 13 01603 g003
Figure 4. Mechanism of urea hydrolysis by urease and mode of action of NBPTO (Adopted from Mazzei et al. [111]).
Figure 4. Mechanism of urea hydrolysis by urease and mode of action of NBPTO (Adopted from Mazzei et al. [111]).
Agronomy 13 01603 g004
Figure 5. (a) Example of a “space-filling” 1,3,4-oxadiazole urease inhibitor with key interactions with neighboring residues. Adopted from Hanif et al. [119]. (b) Location of the mobile flap of urease relative to the active site (top) and a known inhibitor 1,4-benzoquinone which covalently modifies the Cys-containing mobile flap (bottom). Adopted from Mazzei et al. [120].
Figure 5. (a) Example of a “space-filling” 1,3,4-oxadiazole urease inhibitor with key interactions with neighboring residues. Adopted from Hanif et al. [119]. (b) Location of the mobile flap of urease relative to the active site (top) and a known inhibitor 1,4-benzoquinone which covalently modifies the Cys-containing mobile flap (bottom). Adopted from Mazzei et al. [120].
Agronomy 13 01603 g005
Figure 6. Exemplary substrates oxidized by the ammonia monooxygenase (AMO), adopted from Bédard and Knowles [133], Hubley et al. [135], Hyman and Wood [136], Hyman and Wood [137], Stirling and Dalton [138], Voysey and Wood [139], Ward [140], and Hyman et al. [141]. AMO catalyzes the oxidation of various structural motifs including ammonia, methane, ethylene, cyclohexane, and benzene, which releases 2 electrons (2 e). Simultaneously, O2 is reduced, which requires 4 e. The missing 2 e are provided from the set of electrons released during the subsequent oxidation processes catalyzed by HAO. The schematic overview is shown in Figure 7.
Figure 6. Exemplary substrates oxidized by the ammonia monooxygenase (AMO), adopted from Bédard and Knowles [133], Hubley et al. [135], Hyman and Wood [136], Hyman and Wood [137], Stirling and Dalton [138], Voysey and Wood [139], Ward [140], and Hyman et al. [141]. AMO catalyzes the oxidation of various structural motifs including ammonia, methane, ethylene, cyclohexane, and benzene, which releases 2 electrons (2 e). Simultaneously, O2 is reduced, which requires 4 e. The missing 2 e are provided from the set of electrons released during the subsequent oxidation processes catalyzed by HAO. The schematic overview is shown in Figure 7.
Agronomy 13 01603 g006
Figure 7. Membrane-bound AMO oxidizes NH3 to NH2OH, which is further oxidized to NO2 by HAO. The four electrons generated during these processes are subsequently transferred to cytochrome C554, membrane cytochrome Cm552, and ubiquinone-8 (QH2), from where two electrons are fed into the electron transport chain (ETC) and two electrons are transferred back to the AMO to deliver electrons for the reduction of oxygen. In the presence of nitrification inhibitors, such as DCD, Nitrapyrin, and DMP, the first step of ammonia oxidation is inhibited.
Figure 7. Membrane-bound AMO oxidizes NH3 to NH2OH, which is further oxidized to NO2 by HAO. The four electrons generated during these processes are subsequently transferred to cytochrome C554, membrane cytochrome Cm552, and ubiquinone-8 (QH2), from where two electrons are fed into the electron transport chain (ETC) and two electrons are transferred back to the AMO to deliver electrons for the reduction of oxygen. In the presence of nitrification inhibitors, such as DCD, Nitrapyrin, and DMP, the first step of ammonia oxidation is inhibited.
Agronomy 13 01603 g007
Figure 8. Loss of NH3 through volatilization after application of urea (UR) and urease (NBPT) and nitrification (DCD) inhibitors. Figure taken from Soares et al. [192] with permission from Soil Biology and Biochemistry.
Figure 8. Loss of NH3 through volatilization after application of urea (UR) and urease (NBPT) and nitrification (DCD) inhibitors. Figure taken from Soares et al. [192] with permission from Soil Biology and Biochemistry.
Agronomy 13 01603 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Gupta, S.; Yildirim, S.; Andrikopoulos, B.; Wille, U.; Roessner, U. Deciphering the Interactions in the Root–Soil Nexus Caused by Urease and Nitrification Inhibitors: A Review. Agronomy 2023, 13, 1603. https://doi.org/10.3390/agronomy13061603

AMA Style

Gupta S, Yildirim S, Andrikopoulos B, Wille U, Roessner U. Deciphering the Interactions in the Root–Soil Nexus Caused by Urease and Nitrification Inhibitors: A Review. Agronomy. 2023; 13(6):1603. https://doi.org/10.3390/agronomy13061603

Chicago/Turabian Style

Gupta, Sneha, Sibel Yildirim, Benjamin Andrikopoulos, Uta Wille, and Ute Roessner. 2023. "Deciphering the Interactions in the Root–Soil Nexus Caused by Urease and Nitrification Inhibitors: A Review" Agronomy 13, no. 6: 1603. https://doi.org/10.3390/agronomy13061603

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop