Next Article in Journal
Stalling the Course of Neurodegenerative Diseases: Could Cyanobacteria Constitute a New Approach toward Therapy?
Next Article in Special Issue
Effects of Storage Temperatures on Nitrogen Assimilation and Remobilization during Post-Harvest Senescence of Pak Choi
Previous Article in Journal
The Origin of Discrepancies between Predictions and Annotations in Intrinsically Disordered Proteins
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Nitrogen Journey in Plants: From Uptake to Metabolism, Stress Response, and Microbe Interaction

1
Department of Botany and Plant Sciences, University of California Riverside, Riverside, CA 9250, USA
2
Genetics Department, Faculty of Agriculture, Menoufia University, Shebin El-Kom 32511, Egypt
3
Department of Plant Agriculture, University of Guelph, 50 Stone Road East, Guelph, ON N1G 2W1, Canada
4
Botany Department, Faculty of Agriculture, Menoufia University, Shebin El-Kom 32511, Egypt
5
Maryout Research Station, Genetic Resources Department, Desert Research Center, 1 Mathaf El-Matarya St., El-Matareya, Cairo 11753, Egypt
6
Botany and Microbiology Department, Faculty of Science, Kafrelsheikh University, Kafrelsheikh 33516, Egypt
7
Department of Plant Science, University of Manitoba, Winnipeg, MB R3T 2N2, Canada
8
Department of Agricultural Biochemistry, Faculty of Agriculture, Menoufia University, Shibin El-Kom 32514, Egypt
9
Agriculture and Agri-Food Canada, 1391 Sandford Street, London, ON N5V 4T3, Canada
10
Department of Microbiology and Immunology, The University of Western Ontario, 1151 Richmond Street, London, ON N6A 5B7, Canada
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomolecules 2023, 13(10), 1443; https://doi.org/10.3390/biom13101443
Submission received: 21 August 2023 / Revised: 19 September 2023 / Accepted: 19 September 2023 / Published: 25 September 2023
(This article belongs to the Special Issue Nitrogen Signaling, Transport, and Function in Plants)

Abstract

:
Plants uptake and assimilate nitrogen from the soil in the form of nitrate, ammonium ions, and available amino acids from organic sources. Plant nitrate and ammonium transporters are responsible for nitrate and ammonium translocation from the soil into the roots. The unique structure of these transporters determines the specificity of each transporter, and structural analyses reveal the mechanisms by which these transporters function. Following absorption, the nitrogen metabolism pathway incorporates the nitrogen into organic compounds via glutamine synthetase and glutamate synthase that convert ammonium ions into glutamine and glutamate. Different isoforms of glutamine synthetase and glutamate synthase exist, enabling plants to fine-tune nitrogen metabolism based on environmental cues. Under stressful conditions, nitric oxide has been found to enhance plant survival under drought stress. Furthermore, the interaction between salinity stress and nitrogen availability in plants has been studied, with nitric oxide identified as a potential mediator of responses to salt stress. Conversely, excessive use of nitrate fertilizers can lead to health and environmental issues. Therefore, alternative strategies, such as establishing nitrogen fixation in plants through diazotrophic microbiota, have been explored to reduce reliance on synthetic fertilizers. Ultimately, genomics can identify new genes related to nitrogen fixation, which could be harnessed to improve plant productivity.

1. Introduction

Soil erosion, compaction, acidification, and contamination are significant factors causing soil degradation, affecting 50% of agricultural land and food production [1,2,3]. Consequently, 75 billion tons of fertile soil are lost worldwide due to degradation annually [4]. Moreover, massive amounts of synthetic agro inputs to fertilize crops used by farmers harm environmental health [5,6]. In this sense, applying nitrogen (N) fertilizers has increased sharply by 7.4 times compared with crop productivity, which has only increased by 2.4 times, indicating that crops have been reduced in their ability to use N efficiently [7], increasing food insecurity [8,9]. Therefore, innovative and sustainable strategies in the agro-biotechnological sector have received more attention from botanical scientists worldwide over the last century to solve food security challenges and develop environmentally friendly products [10]. Throughout time, microbial diversity and biomass have been assessed as partners in the plant–microbe association in maintaining soil multifunctionality [11].
N is a vital macronutrient for plants and a crucial component of amino acids that serve as the building blocks of enzymes and proteins in plants. Additionally, N is a part of the chlorophyll molecule, an essential factor in photosynthesis for absorbing sunlight energy, promoting plant growth and grain yield. Even the roots contain this essential element because proteins and enzymes control water and nutrient intake [12]. Recently, Zhang et al. (2022) found that N can significantly impact flowering time [13]. In addition to being a crucial nutrient for plant growth and development, N has a strong correlation with several abiotic stress responses [14]. Moreover, N has a leading role in plant adaptation to the deficiency of macro- and micronutrients. For macronutrients, the processes of N and phosphorus (P) uptake interact together, creating a nutritional balance under fluctuating availability of nutrients [15], while potassium (K) and nitrate (NO3) translocation and intake are positively connected as well [16]. The plant nitrate transporters (NRTs) regulate the K translocation [17]. Regarding micronutrients, ammonium (NH4+) application promotes iron (Fe) uptake [18]; in contrast, NO3 triggers the chlorosis symptoms of Fe deficiency [19]. Sulfur (S), a crucial component of enzymes involved in N metabolism [20], is correlated with N uptake [21]. Alternatively, a reverse relationship between molybdenum (M) and N has been noticed [22]. In addition, through the mechanism of anion-cation balance, approximately 90% of stress responses are influenced by NO3 uptake, which promotes the synergetic transfer of cations, including K+, sodium (Na+), cadmium (Cd2+), and zinc (Zn2+) while inhibiting the absorption of anions like chloride (Cl) and sulfate (SO4) [14]. Under drought stress, NO3 and NH4+ concentrations have unique impacts on plant performance. While the usage of NH4+ reduces the effects of drought on plant development, NO3 application has a reverse influence [23].
In this review, we aim to provide a detailed story of the N journey starting from its uptake from soil either in NO3 or NH4+ form, followed by a structural investigation of the transporters of these two forms. Upon reaching their destination, NO3 and NH4+ are exposed to various metabolic pathways (e.g., oxidation and denitrification). Next, N assimilation occurs under the intersection with glutamine synthetase (GS) and glutamate synthase or glutamine 2-oxoglutarate aminotransferase (GOGAT). Then, a comprehensive survey of the role of N in plant adaptation/tolerance to abiotic stresses, either drought or stress, is discussed. Finally, the detrimental effects of nitrogen mineral fertilization and the possibility of microbial application and genomic approaches are summarized.

2. Nitrogen Uptake

N is an essential macronutrient for plants, and it can be acquired from the soil in inorganic forms (NO3 and NH4+), which are transported across the root plasma membrane by different families of transporters. Under normal soil conditions, N is mainly available in the form of NO3, and four families of transporters mediate NO3 uptake, namely NRT1, NRT2, chloride channel (CLC-1), and slow anion channel-associated 1 homolog 3 (SLAC1/SLAH). These transporters have distinct characteristics in affinity, capacity, regulation, and localization [24]. NH4+ uptake is facilitated by ammonium transporters (AMTs), high-affinity transporters expressed mainly in the root hairs and epidermis [14]. NH4+ is the dominant form of N in flooded or acidic soils, and AMTs-mediated acquisition is crucial for the N demand of plants growing in such conditions [25]. Besides inorganic N, plants can also absorb organic N in the form of amino acids (AAs), which are abundant in soils that receive organic amendments, such as manure or compost [26]. Several AAs transporters have been identified in plant roots, including amino acid permease (AAP1 and AAP5), proline transporter (ProT2), and lysine and histidine transporters (LHT1 and LHT6) [27,28]. These transporters have different substrate specificities and expression patterns contributing to the uptake of a range of amino acids from the soil solution [27,29]

3. Nitrogen Transporters

3.1. Plant Nitrate Transporters

Plants acquire NO3 from the soil through various transporters that belong to different families [30]. The NRT1 and NRT2 families are the primary transporters involved in NO3 uptake, and they have distinctive roles in plant growth and seed development. The NRT1 family is a large and diverse group of nitrate peptide transporter family (NPF). There are 53 NRT1 genes in Arabidopsis and 93 in rice [31]. These genes can be classified into 8–10 subfamilies; NRT1.1 and NRT1.2 are responsible for NO3 root uptake. The NRT1 family can transport NO3 and other substrates such as hormones, nitrite, amino acids, peptides, chloride, glucosinolates, and jasmonate-isoleucine [32]. The NRT2 family is a smaller, more specific group of transporters expressed mainly under low NO3 conditions. In Arabidopsis, four NRT2 transporters (NRT2.1, NRT2.2, NRT2.4, and NRT2.5) function in nitrate influx and account for 95% of nitrate uptake under low NO3 concentrations [33]. However, NRT2.1 and NRT2.2 are significant members of the NRT2 family for nitrate uptake. The NRT2 family has different spatial and temporal expression patterns, and they transport NO3 from different sources to different tissues. For example, NRT2.4 and NRT2.5 absorb NO3 from the soil root hairs, while NRT2.1 and NRT2.2 transport it from the apoplast to the cortex and endodermis cells. Moreover, NRT2.5 is induced by long-term starvation and enhances NO3 uptake from both shoots and roots of mature plants [33,34]. Furthermore, another transporter family called NRT3 is also involved in NO3 transport in plants. The NRT3 family consists of two members: NRT3.1 and NRT3.2. These transporters form a complex with NRT2 transporters and regulate their activity and stability [35].
NO3 transceptor (transporter/receptor) NRT1.1 regulates the expression levels of many NO3 assimilation pathway genes by sensing the external NO3 concentration and modulating the root growth accordingly [36,37]. NO3 transporters are regulated through phosphorylation, mediated by calcium-dependent calcineurin B-like (CBL) and CBL-interacting protein kinase (CIPK) [38,39]. The activity and specificity of NRT1.1 are regulated by two CIPKs, CIPK8 and CIPK23, which interact with NRT1.1 and mediate low- and high-affinity responses, respectively [40]. At high NO3 concentrations, NRT1.1 activates the expression of Arabidopsis nitrate regulated 1 (ANR1), a transcription factor that promotes lateral root growth and proliferation [41]. At low NO3 concentrations, NRT1.1 inhibits lateral root development by controlling auxin levels and meristem activation [42]. To transmit the NO3 signals from NRT1.1 to the nucleus, calcium acts as a secondary messenger that modulates the expression of NO3-responsive genes. Three calcium-dependent protein kinases (CPK10, CPK30, and CPK32) and their partner CBLs are involved in this process [43].
Several NRTs are involved in the loading and unloading of NO3 from the xylem and phloem, which affects the distribution and availability of NO3 within the plant. For example, NRT1.5, NRT1.8, and NRT1.9 are expressed in the xylem and phloem, and they function in the influx/efflux, removal, and loading of NO3 from/to the root vascular tissues, respectively [43,44]. Other studies have shown that overexpressing NRTs in different plant species can improve N-use efficiency (NUE) as well as plant biomass to nitrogen input. However, the effects of NRT overexpression may depend on the tissue specificity, NO3 concentration, and interaction with other genes. For instance, only NRT1 transporters have been successfully overexpressed in both leaves and roots to enhance NUE, while the role of NRT2 transporters in this process is still unclear [45]. Recently, a chimeric NRT1 transporter, AtNC4N, was overexpressed in the phloem of old leaves in Arabidopsis, rice, and tobacco, and it increased NO3 uptake and NUE under low nitrate conditions [46]. Another NRT1 transporter, OsNRT1.1A (OsNPF6.3), was found to improve NUE, flowering time, yield, and maturity in rice [47]. Several other NRTs, such as OsNRT1.1B (OsNPF6.3B), OsNRT2.1, OsNRT2.3a/b, OsPTR9, OsAMT1.1, and qNGR9, have also been reported to enhance NUE under high or low nitrate levels in rice [47,48,49,50].
Moreover, some genes that regulate or interact with NRTs have been identified as potential targets for improving NUE in plants. For example, the transcription factor OsNAC42 activates the expression of several NRTs and increases nitrate assimilation and NUE in rice. The nitrate reductase gene OsNR2 converts nitrate to ammonium, affecting the expression of NRTs and other nitrogen-related genes in rice. The nitrate transporter OsNPF4.5 is regulated by a microRNA (miR827) and modulates nitrate distribution and remobilization in rice. These studies demonstrate the complexity and diversity of nitrate transport and signaling in plants as well as provide valuable insights for improving NUE and crop production [51]. Notably, NUE can be improved by transcription factors (TFs) in multiple crops [52,53]. TFs are primarily proteins (rarely RNA) that control the transcription of genes by interacting with their unique promotor sequences [54]. Numerous TF families have been found and proven to interact with particular DNA sequences on regulatory regions of NO3-responsive genes. According to reports, these TF families act as essential regulators of the plant N response, which encompasses MADS-box [55], NLP [53], b-ZIP [56], LBD [57], NAC [58], MYB [59], GARP [60], TCP [61], AP2/ERF [62], and zinc/finger proteins [63].
NRT1, NRT2, and NRT3 have different structural conformations (Figure 1). As a result, the Pfam database collects these genes under distinct families: PF00854 for NRT1, PF07690 for NRT2, and PF16974 for NRT3 [64]. PF00854 refers to the proton-dependent oligopeptide transporter (POT) family (also known as the peptide transporter (PTR) family) and is a class of energy-dependent transporters discovered in organisms ranging from bacteria to humans [65]. However, some family members are NO3 permeases, while others are in histidine transport [66]. The NRT1 transporter adopts a standard major facilitator superfamily (MFS) fold, which is described by 12 transmembrane helices (TMHs) with a pseudo 2-fold axis connecting the N-terminal (TMH1–6) and C-terminal (TMH7–12) domains [67].
NRT2 members belong to the PF07690 (MSF) family, representing the largest family of secondary transporters, with members ranging from Archaea to Homo sapiens. MFS proteins target a diverse range of substrates in both directions across the membrane, including ions, carbohydrates, lipids, amino acids and peptides, nucleosides, and other small molecules, in many cases catalyzing active transport by converting the energy stored in a proton electrochemical gradient into a concentration gradient of the substrate [68]. NRT3 candidates lie under the PF16974 family, which contains a C-terminal transmembrane domain required for high-affinity nitrate absorption [69,70]. NRT3 is implicated in the inhibition of lateral root initiation in the presence of high sucrose-to-nitrogen ratios in the medium. As a result, the NRT2 and NRT3 component systems are believed to be implicated in the signaling pathway that integrates nutritional inputs for the regulation of lateral root architecture [71]. In Arabidopsis, the functional unit of the high-affinity nitrate influx complex is most likely a tetramer composed of two NRT2 and NRT3 subunits each [72].

3.2. Plant Ammonium Transporters

As mentioned earlier, NO3 and NH4+ are the principal inorganic nitrogen sources absorbed by plant roots [73,74]. Because NH4+ requires less energy than NO3 assimilation, NH4+ is the preferred form of N uptake when plants grow under N deficiency [75,76]. Although NH4+ uptake is more efficient, high absorbed amounts of NH4+ by the plant can be hazardous [77,78,79]; thus, the NH4+ uptake system in plants receives special attention [80,81,82]. Recent research indicates that AMTs also play a role in a variety of other physiological processes such as transporting NH4+ from symbiotic fungi to plants [83,84], delivering NH4+ from roots to shoots [85], transmitting NH4+ in leaves and reproductive organs [86,87,88], and encouraging resistance to plant diseases through NH4+ transport [89,90]. Aside from being a transporter, many AMTs are essential for root growth in the presence of NH4+ [81,91,92,93]. To avoid the adverse effects of insufficient or excessive NH4+ intake on plant growth and development, AMTs activities are fine-tuned not only at the transcriptional level through the involvement of at least four transcription factors [94,95,96,97] but also at the protein level through phosphorylation [98,99], pH [100], and endocytosis [101].
The NH4+ transporter domain (PF00909) comprises two structural copies of five helices plus one extra C-terminal helix. It has been defined as an 11-times-spanning membrane channel. Ideal NH4+ transporter domain is composed of 11–12 transmembrane spots, with feature sequences “D (F Y W S) A G (G S C) X2 (L I V) (E H) X2 (G A S) (G A) X2 (G A S) (L F)” at transmembrane spot 5 and “D D X (L I V M F C) (E D G A) (L I V AC) X3 H (G A L I V) X2 (G S) X (L I V A W) G” at transmembrane region 10 [81]. AMTs can be divided into two subtypes: AMT1 and AMT2 or methylammonium permeases (MEP) [81,85,102]. Figure 1 illustrates the diversity and complexity of the N transporters in plants and their distinctive features.

3.3. Plant Nitrite Transporters

Cytosol is the first cellular compartment to encounter NO3 in the root that needs to be converted to NO2 to make organic compounds such as amino acids. The cytosolic NO2 is toxic to the cytosol and needs to be transported to the chloroplasts, where it is reduced to NH3 by NIR [103]. Cytosolic NO2 needs to be actively transported across the membrane of the chloroplasts by a specific transporter to avoid accumulation and toxicity. Several types of plastids and plasma membrane-localized NIR have been described in Escherichia coli, Cyanobacteria, green algae, red algae, Aspergillus nidulans, diatoms, and a haptophyte, but there was no evidence of a nitrite-specific plastid and plasma membrane transporter in higher plants [103,104]. Consequently, it has been hypothesized that NO3 and NO2 enter the plasma membrane via similar transporters due to previously observed competitive reduction of NO2 absorption by NO3 and vice versa [105].
In 1988, Brunswick and Cresswell found that illuminated chloroplasts take up NO2 in vitro [104]. In 2007, Saguira et al. revealed that CsNitr1-L is the NO2 transporter in the chloroplasts of plants and belongs to the proton-dependent oligopeptide transporter (POT) family [106]. CsNitr1-L, a homologue of NRT1 transporter, can transport NO2 depending on the energy availability. In 2013, Kotur et al. used 13NO2 to characterize nitrite influx and nitrite-specific transporter into roots of Arabidopsis thaliana [105]. Fortunately, they successfully confirmed that Arabidopsis roots include a transporter specific for nitrite [105]. Furthermore, in 2014, Maeda et al. succeeded in characterizing two genes, AT5G62720 (AtNITR2;1) and AT3G47980 (AtNITR2;2), which encode an integral membrane protein related to the NO2 transport activity from Arabidopsis, and these two genes were detected based on homologs genes from cyanobacterial genomes [107].
AtNITR2;1 and AtNITR2;2 are HPP-domain-containing proteins (Figure 1). These two transporters belong to PF04982, constituting integral membrane proteins with four transmembrane-spanning helices [108]. Interestingly, their work has great value, as the HPP proteins had no functional annotation before their work [107]. Moreover, the kinetic analyses showed that the proteins that encode from these previous two genes of Arabidopsis have a much higher affinity for NO2 than the cyanobacterial proteins [107]. AtNITR2;1, isolated from the chloroplasts mutant lines, showed much lower NO2 uptake activity than the chloroplasts isolated from the wild-type Col-0 plants. AtNITR2;2, expressed in roots and not detectable in shoots, had a putative transit peptide similar to that of AtNITR2;1 and demonstrated low but significant activity of NO2 transport in the cyanobacterial cell. Finally, it was discovered that NO2 increased the expression of AtNITR2;1 and AtNITR2;2 under the direction of NIN-like proteins, indicating that these two proteins are nitrate-inducible parts of the nitrite transport system of plastids.

4. Nitrogen Metabolism

N is among the most widely distributed elements in the lithosphere, atmosphere, hydrosphere, and biosphere [109]. The lithosphere contains 94% of all nitrogen on Earth (e.g., NO3, NO2, and NH4+), with the remaining 6% in the atmosphere (e.g., NO, N2O, and N2) and a trace (0.006%) in the hydrosphere and biosphere (e.g., NO3, NO2, and NH4+). N is the fourth most prevalent element in the biosphere after oxygen, carbon, and hydrogen and is an essential component of total biomass [110,111]. The relative quantity of N in the biosphere reflects the relevance of N to living organisms. Although organic N, such as amino acids, hormones, and nucleic acids, is vital to life on Earth [112,113,114,115], The global N cycle is dominated by inorganic N molecules (i.e., NO3, NO2, nitrous oxide (N2O), NO, dinitrogen (N2), and NH4+) [112,116,117,118]. Figure 2 summarizes all N metabolism pathways.
NO3 is the most oxidized form of N, while NH4+ is the most reduced form. Many microorganisms, mainly microbes, drive the global N cycle by metabolizing N through a variety of redox processes for energy transduction, detoxification, or assimilation via various sub-pathways such as N fixation (KEGG;M00175), assimilatory NO3 reduction (KEGG;M00531), dissimilatory NO3 reduction (KEGG;M00530), denitrification (KEGG;M00529), nitrification (KEGG;M00528), complete nitrification, complete ammonia oxidation (comammox) (KEGG;M00804), and anammox (KEGG;M00973).
Moreover, each sub-pathway from these previous sub-pathways consists of a list of genes and their encoded enzymes that are involved in the various biogenic processes in the nitrogen cycle. The nitrogen fixation path consists of nitrogenase molybdenum-iron protein alpha chain (NIFD), vanadium-dependent nitrogenase alpha chain (VNFD), and nitrogenase delta subunit (ANFG) genes [119,120,121].
In this sub-pathway, these previous genes were used to convert N2 to NH3 by oxidoreductase reaction. NH3 is widely used in different vital pathways for living organisms, such as arginine biosynthesis (KEGG;map00220), purine metabolism (KEGG;map00230), alanine, aspartate, and glutamate metabolism (KEGG;map00250). Also, NH3 is involved in glycine, serine, and threonine metabolism (KEGG;map00260) as well as cyanoamino acid metabolism (KEGG;map00460), glyoxylate and dicarboxylate metabolism (KEGG;map00630), and lipoic acid metabolism (KEGG;map00785). NH3 is a component in biosynthesis of secondary metabolites (KEGG;map01110), microbial metabolism in diverse environments (KEGG;map01120), carbon metabolism (KEGG;map01200), biosynthesis of amino acids (KEGG;map01230), proximal tubule bicarbonate reclamation (KEGG;map04964), collecting duct acid secretion (KEGG;map04966), protein digestion and absorption (KEGG;map04974), Vibrio cholerae infection (KEGG;map05110), and epithelial cell signaling in Helicobacter pylori infection (KEGG;map05120) [119,120,121].
In addition, the anammox path consists of hydrazine dehydrogenase (HDH), hydrazine synthase subunit (K20932), and nitrite reductase (NO forming) (NIRK) genes, and these previous genes can convert NO2 to N2 or NH3 using nitrite reductase enzyme [122,123]. In that context, the denitrification path consists of four genes: nitrous-oxide reductase (NOSZ), NIRK, nitrate reductase (cytochrome) (NAPA), and nitric oxide reductase subunit B (NORB) [124,125,126]. In this sub-pathway, the denitrification enzymes convert NO3 to N2 through NO2, NO, and N2O. Furthermore, assimilatory and dissimilatory NO3 reduction paths have different genes, such as nitrate reductase/nitrite oxidoreductase, alpha subunit (NARG, NARZ, and NXRA), NAPA, nitrite reductase (NADH) large subunit (NIRB), nitrite reductase (cytochrome C-552) (NRFA), ferredoxin-nitrate reductase (NARB), nitrate reductase (NAD-(P)-H) (NR), assimilatory nitrate reductase catalytic subunit (NASC and NASA), nitrite reductase (NAD-(P)-H) (NIT-6), ferredoxin-nitrite reductase (NIRA), and nitrite reductase (NAD(P)H) large subunit (NASD and NASB) that can convert NO2 to NH3 [127,128,129,130]. Figure 2 and Table 1 show the enzymes and genes involved in the N cycle.

4.1. GS and GOGAT

N uptake for plant growth has been extensively studied and discussed previously [172]. N uptake involves an active transport across the plasma membrane of both root epidermal and cortical cells. The gene expression patterns, post-translational regulation, and the localization of NRTs and AMTs enable directed transport from the epidermis to vascular tissue and acclimation to various concentrations of N in the soil [173]. Mostly, the plant uptakes N from the soil in the form of NO3 that is reduced by NR to NO2, which is then imported into the chloroplast and reduced further by NIR into NH4+. NO2 reduction to NH4+ occurs in the stroma of chloroplasts and is catalyzed by the enzyme NIR. NIR reduces NO3 to NO2 using electrons from NAD(P)H, and the reduced ferredoxin (Fdred) generated due to the photosynthetic electron transport is the electron donor for NO2 reduction. NH4+ is combined with glutamate to be assimilated into glutamine [174]. NR is regulated by NO3, N, light, growth conditions, hormones, N metabolites, and phosphorylation [175,176,177,178].
In various biological systems, the interconversion of glutamate and glutamine plays a central role in N assimilation, transport, and recycling. GS and GOGAT are vital enzymes orchestrating these interconversions, ensuring optimal N utilization and maintaining N homeostasis [112]. The roots are the principal site of NH4+ assimilation, where the enzyme GS facilitates this conversion [179,180]. Following GS action, one of the amino groups from glutamine is transferred to 2-oxoglutarate (2-OG) to synthesize glutamate, and this reaction is catalyzed by GOGAT (Figure 3). The coordinated action of GS and GOGAT is responsible for these interconnected processes [181,182,183]. Experimental studies involving reverse genetic analysis have demonstrated that the absence of genes encoding NH4+-responsive GS/GOGAT isoenzymes reduces NH4+ assimilation, particularly in the roots, resulting in impaired growth [184,185,186,187]. These findings suggest that the NH4+-responsive forms of GS/GOGAT have a central role in the primary assimilation of NH4+ within the roots [188]. As the expression of these isoenzyme-encoding genes increases with NH4+ availability, it is plausible to infer that plants possess a regulatory transcriptional network that modulates their gene expression in response to NH4+ availability [185].
Plant GS (EC6.3.1.2), a highly conserved enzyme, catalyzes glutamine synthesis from glutamate and ammonium ions, conserving the N atom in an organic molecule. This conversion is fueled by the energy released through the hydrolysis of ATP to ADP, making GS a crucial component of N metabolism across different life forms [189]. In addition, the GOGAT in synthesizing glutamate involves transferring the amide amino group of glutamine to 2-OG through a reductant-driven process, yielding two glutamate molecules [190]. Within the realm of plants, this enzyme exists in two distinct variants: one produces the ferredoxin (Fd) as the electron donor (EC 1.4.7.1), while the other employs NADH as the electron donor (EC 1.4.1.14). The Fd-dependent form of the enzyme is notably abundant within the chloroplasts of photosynthetic tissues. In these regions, it utilizes light energy directly to serve as a reductant source. Figure 3 provides an overview of the N cycle, transport, and utilization in plants.
Conversely, the NADH-dependent variant, also present within plastids, is primarily concentrated in non-photosynthesizing cells. The reductant is sourced from the pentose phosphate pathway [191]. Notably, the expression of the Fd- and NADH-dependent forms of glutamate synthase seems to exhibit dissimilar patterns across distinct plant tissues. It has been suggested that in most plants, two genes encode each form of GOGAT [190].

4.2. Isoforms of GS and GOGAT

According to subcellular distribution patterns, investigations into plant GS enzymes have identified two main categories. The initial group, designated as GS1, is primarily situated within the confines of the cytoplasm. Conversely, the second group, GS2, is predominantly found within plastid structures [192,193]. Various multigenic families are responsible for encoding multiple isoforms of GS1, whereas the plastidial GS2 arises from a limited number of nuclear genes [194]. Typically, GS2 is linked to the process of (re)assimilation of NH4+ in leaves, whereas GS1 is associated with recycling N within the plant. However, the relative activity levels exhibited by GS1 and GS2 are subject to variation based on species, specific plant organs, N sources, developmental stages, and prevailing environmental conditions, including abiotic stress factors. This variability underscores the intricate involvement of these isozymes, reflecting a multifaceted role [193,195]. In wheat research, a comprehensive exploration has unveiled four distinct variations of GS genes: GS1, GS2, GSr, and GSe [192]. Employing advanced high-performance liquid chromatography (HPLC) methods, ref. [196] successfully separated GS and discovered two subtypes in leaves: cytoplasmic subtype GS1 and chloroplast subtype GS2. Both subtypes’ content and activity fluctuate in tandem with developmental processes. The subcellular localizations of GSr and GSe align with the cytoplasmic compartment [197]. In addition, the advancement of genome sequencing has unveiled various isoenzymes of GS/GOGAT present in plants. Among these, specific isoenzymes such as GS1,2 and NADH-GOGAT1 from rice [198] as well as GLN1,2 and NADH-GOGAT (GLT1) from Arabidopsis [185,199] are expressed in plant roots in response to NH4+ supply. These isoenzymes’ expression at transcript and protein levels highlights their specificity in NH4+ assimilation.

4.3. Nitrogen Metabolism in Microalgae

Microalgae are unicellular aquatic microorganisms with sizes between 1 µm and 2 mm. They can generate biomass using light and carbon dioxide through the photosynthetic process. Due to their remarkable tolerance for environmental stress, they can be grown in various environmental conditions, including freshwater, saltwater, or wastewater [200]. After C, N is quantitatively the most significant element contributing to the dry matter of microalgal cells, accounting for between 1 and 10% of their dry weight. C and N metabolism are related in microalgae since they share the following: (a) C provided directly by respiration of CO2 or assimilated organic C and (b) the energy obtained by the TCA cycle and mitochondrial electron transport chain. To create the amino acids glutamate, glutamine, and aspartate, the primary absorption of inorganic N (NH4+) needs carbon skeletons in the form of keto acids (2-OG and oxaloacetate) as well as energy in the form of ATP and NADPH [201]. There are four fundamental phases in the NO3 assimilation mechanism in eukaryotic microalgae and plant leaves: NAD (P) H-nitrate reductase catalyzes NO3 reduction into NO2 in the cytosol of cells, which is required for (a) NO3 transport into the green cell, and (b) NO3 reduction into NO2 in the cytosol of cells. In addition, in Cyanobacteria, ferredoxin-nitrate reductase catalyzes the reduction of NO3 to NO2, (c) NO2 transport into the chloroplast of cells via NAR1 family and its subsequent reduction into NH4+ in a six-electron reaction catalyzed by ferredoxin-nitrite reductase, and (d) the process mediated by the ATP-dependent GS- GOGAT cycle, which results in the formation of L-glutamate by the incorporation of NH4+ into the carbon skeleton of 2-OG [103]. Microalgae mostly use the GS-GOGAT pathway to assimilate NH4+ as well as via the amination of 2-OG catalyzed by NAD(P)H glutamate dehydrogenase [NAD(P)H-GDH [202,203].

5. The Role of Nitrogen in Plant Response to Abiotic Stress

Plants must endure abiotic stressors, including drought, salt, and severe temperatures, because they are immobile. The spread of plants is significantly constrained by these stresses, which also change plant growth and development and lower agricultural output [204]. Since N is considered the most crucial nutrient for plant growth quantitatively, plants have developed effective methods to control N levels in response to complicated stresses [14].

5.1. Nitrogen and Drought Stress

Drought stress significantly damages crops worldwide, particularly in arid and semi-arid regions [205]. Hence, plants have evolved to adapt their resilience against drought via avoidance (stomata adjustment to reduce transpiration rate) and tolerance (antioxidants and osmolytes production) of drought [206] (Figure 4). Remarkably, the concentrations of NO3 and NH4+ have impacts on plants’ underwater limitations. The application of NH4+ alleviates the effect of drought on plant development, while NO3 has the opposite impact [54]. In that sense, N fertilizers might enhance the physiological responses of plants under drought stress by elevating the concentrations of N and chlorophyll and promoting PSII photochemical activity [207]. It has been shown that adding N (NH4NO3) improved water-use efficiency (WUE) by increasing plant dry mass and decreasing water loss [206].
Additionally, N may lessen drought’s inhibitory effects on photosynthesis and prevent C starvation [208]. On the other hand, low N availability can make plants more sensitive to dehydration and cause protein degradation, which lowers the amount of N-containing osmolytes like proline [206,208]. N supply enhances osmotic adjustment and the activity of antioxidant enzymes via proline accumulation induction to mitigate the physiological harm from drought stress. However, the N effects depend on the ionic form of N supplied [209]. NH4+ is essential in promoting plant drought tolerance by increasing root numbers (surface area) and subsequently improving and facilitating water uptake. Also, decreased aerenchyma development may contribute to NH4+-promoted drought tolerance [14]. The influence of NO3 on drought responses is linked to NO3 transport/assimilation in plants. Numerous NO3 transport/assimilation-related genes, such as NRT2.5, GS, and GOGAT, are repressed in response to drought stress [210]. Interestingly, a high N concentration enhances root growth and photosynthetic traits, alleviates drought stress via stomatal conductance, and increases the photosynthetic rate [211]. Plant nutrient management is crucial to promote abiotic stress tolerance in cotton plants [212]. The use of nitrogen fertilizer either in the soil or foliar spray can decrease the toxicity of reactive oxygen species (ROS) by increasing the peroxidase (POD), catalase (CAT), and superoxide dismutase (SOD) levels in plant cells [205,212]. These enzymes improve the photosynthetic rate, promoting stress tolerance by scavenging ROS and reducing photooxidation to preserve chloroplast membrane integrity [212,213,214]. Applying nitrogen fertilizers (N60 kg.ha−1) can increase the tolerance of drought stress in wheat plants combined with the highest yield [205].
Under drought stress, GS1 and GS2 exhibited differential expression among different plant organs and species, with a prominent increase of GS1 in roots and leaves, while the total GS activity showed a decrease in potato root [215,216]. A comparative proteomic analysis under drought stress conditions in wheat was studied [217]. The researchers found that drought-tolerant wheat varieties exhibited a significant upregulation of GS2 expression, indicating its potential role in enhancing drought tolerance. Similar results were obtained by Nagy et al. [218], who explored GS activity in wheat leaves and observed that older leaves had lower GS activity than flag leaves in drought-tolerant and sensitive varieties under normal conditions. During drought stress, GS activity in sensitive varieties’ flag leaves decreased compared to tolerant varieties. This suggested that GS activity could indicate drought stress sensitivity and tolerance, with elevated cytoplasmic-to-plastid GS ratio and premature senescence as markers. Exposure to heavy metals such as cadmium, molybdenum, copper, and aluminum led to the down-regulation of GS and decreased its total activity in different plant species, including rice, soybean, tomato, and tea [219,220,221,222].
Like all organisms, plants must respond to various extreme environmental cues by responding to various internal signals. NO is a small lipophilic molecule that diffuses through plant cell membranes as active signals to thrive and survive under those conditions. In addition, NO is a relevant signal molecule in many plant processes. Several pathways for NO production, either oxidative in the presence of O2 or reduction from NO2, have been described in plants [223]. NO regulates gene expression, modulates enzyme activities, and acts as a metabolic intermediate in energy regeneration. However, the mechanisms and proteins responsible for its synthesis remain complex, with many unresolved questions.
Moreover, several studies showed the involvement of NO in reproductive processes, control of development, and the regulation of physiological responses such as stomatal closure. NO also regulates the expression of several genes involved in the synthesis of and response to pathogen attacks as well as reproductive mechanisms that operate during pollen recognition by the stigma [224,225,226,227]. NO is a high redox to reactive N species (RNS) formed in living cells under normal and stressful biotic and abiotic conditions, and NO also functions as a detoxifier and lessens adverse effects when a plant’s ROS content reaches dangerous levels [228]. It was reported that the synthesis of NO increased in plant cells under drought stress conditions, and it has been suggested that this may help enhance plants’ acclimation responses to cope with water limitation [229]. Drought stress enhanced NO synthesis in beat, pea, tobacco, cucumber, grapevine, and rice plants; likewise, the activity of NO synthase (NOS)-like and NO release rate increased under dehydration conditions [230]. NO has a vital role in respiratory function, called pathways of electron transport in mitochondria; it adjusts ROS as well as activates defense strategies via promoted antioxidant formation in plants under various abiotic stresses [228]. NO plays a crucial role in mediating hormonal activities and influencing gene expression as well as protein activity during sensing and signaling processes that enable plants to survive under environmental stressors. The exogenous supply of NO increased drought tolerance in various plants such as maize, rice, and cucumber due to the increase of antioxidant defense, which acts to scavenge ROS, improve cellular membrane stability, and preserve photosynthesis as well as water status [230,231].
Drought induces the synthesis of abscisic acid (ABA), which regulates several critical plant processes like seed germination and stomatal conductance, with the help of signaling molecules like NO and ROS [232]. Interestingly, NO acts as a critical signaling molecule promoting ABA-induced stomatal closure in case of hyperosmotic stress [233]. The plants’ main enzyme responsible for NO production is the molybdoenzyme NR, which contains two subunits, each bearing three prosthetic groups: FAD, heme b557, and molybdenum. In an NR subunit, molybdenum is bound to a tricyclic pyranopterin and chelated by a dithiolene named the molybdenum cofactor (Moco). Amongst the reducing sources of NO in plants, NO2 can be reduced to NO by a family of five molybdenum (Mo)-containing enzymes, which include NR, xanthine oxidase reductase (XOR), aldehyde oxidase (AO), sulfite oxidase (SO), and mARC [234,235].
In a study by Berger et al. (2020) three NR genes (MtNR1, MtNR2, and MtNR3) were discovered in the genome of Medicago truncatula, and they study addressed their expression, activity, and potential involvement in NO production during the symbiosis between M. truncatula and Sinorhizobium meliloti [235]. These results revealed that MtNR1 and MtNR2 gene expression and activity are correlated with NO production throughout the symbiotic process and that MtNR1 is mainly involved in NO production in mature nodules [235]. However, when plants were grown in soil containing NH4+, the functional role of NR was eliminated [236]. Moreover, NO can alleviate the negative impact of ROS through lipid peroxidation, enhancing photosynthesis as well as the expression of antioxidant enzymes via mitogen-activated protein kinase (MAPK) and other signaling pathways [228]. It was found that exogenous application of NO caused a significant increase in seedling growth of cucumber by decreasing seedlings injury and enhancing the antioxidant activities and chlorophyll and free proline contents under stress conditions [237]. Moreover, an exogenous supply of NO in wheat markedly preserved high relative water content via decreasing transpiration and increasing K content and mitigated cell membrane damage [238]. Figure 4 explains how NO can work as a signaling molecule that can protect plants from drought stress by modulating various physiological and biochemical processes.

5.2. Nitrogen and Salinity Stress

Salinity stress has a destructive impact on the metabolism and productivity of crops worldwide, resulting in the remarkable conversion of agricultural arable soil into unproductive land [239]. The high level of salt ions in agricultural soil may be harmful to plants, causing salt stress, where sodium (Na+) and chlorine (Cl) concentrations increase in saline soil and affect the natural physiological process in plants [240]. N application is an efficient way to improve plants under saline conditions. N is a necessary nutrient that mediates various molecular, physiological, and cellular responses essential to plant survival as well as different signal transduction pathways linked to plant defense systems and salt-stress tolerance [241,242]. Moreover, salt stress affects ammonification and nitrification in the soil because Cl competes with NO3, and NH4+ competes with Na+ and creates ion toxicities and inequalities, which can limit the processes of N uptake, transport, and assimilation [243,244]. Salt stress has inhibited various enzyme activities that contribute to N assimilation in plants such as maize, tomato, rice, cowpea, and mung bean, reducing N uptake and utilization [245].
Interestingly, it was reported that the interaction between salt stress and N had an impact on plant growth, and the effect of N on salt-exposed plants was dependent on the N rate [240]. In the case of maize, excessive N treatment can successfully mitigate the deleterious effects of salt stress [246]. Similar findings in tomatoes showed that supra-optimum N was better than optimum N in reducing salt stress [247]. Under salinity stress, the application of N significantly increased both N and K uptake and decreased Na accumulation in wheat seedlings. Furthermore, N application can protect wheat seedlings in saline conditions via the upregulation of osmolytes, antioxidant system, and secondary metabolite accumulation [239]. Even though excessive N application exacerbated the negative effect of salt stress on wheat and rice, moderate N supply could ameliorate the deleterious effect of salt stress. Therefore, applying an appropriate N rate is the key to mitigating the adverse effects caused by salt stress [240].
On the contrary, it was reported that the application of NO3 resulted in an increase of Na+ uptake and loading of the sodium into the xylem, increasing root inhibition caused by the salinity. Na+ and NO3 co-transportation was suggested based on NO3-dependent Na+ uptake data at various Na+ concentrations [14]. Similarly, NO3-dependent transport mechanisms in salt-stress environments encourage Na+ ion uptake and loading into the xylem, which may represent a main pathway for Na+ accumulation in Arabidopsis shoots [241].
It is worth noting here that GS isoforms imply a multifaceted role for GS in navigating through various dimensions of abiotic stress tolerance [193,248]. For instance, in rice plants exposed to salinity stress, GS1 increased in the root at the seedling stage and old leaves with no change in young leaves, while GS2 was down-regulated in old and young leaves [249].
It has been demonstrated that NO can act as a signal to induce the resistance of salt stress by raising the K:Na ratio in plant cells, which is based on an increase in the H1-ATPase activity of plasma membrane in calluses from maize seedlings, reed (Phragmites communis) plants, and poplar trees [230]. Exogenous NO can improve salt tolerance under salinity conditions by stimulating the K+/Na+ ratio, the Na+/H+ antiport in the tonoplast, and proton pump activity [250]. Under salt stress, NO improves the absorption and translocation of several macro- and micronutrients, including K, Fe, Mg, and Zn, which improve respiration and chlorophyll biosynthesis [251]. The uptake of mineral nutritional elements was improved due to NO application, thereby diminishing the harmful effect of high salinity on strawberry plants [252]. The exogenous NO supply can protect chickpea plants from salt stress-induced oxidative harm by promoting the biosynthesis of photosynthetic pigment and antioxidant enzymes osmolyte accumulation, thereby improving plant development under saline stress [253]. Furthermore, under salt stress conditions, exogenous NO treatment enhanced wheat seed germination by converting starch to sugars, where the application of NO led to an elevation in the effectiveness of α- and β-amylase enzymes [251,254]. Under salt stress, NO treatment boosted plant development and demonstrated a protective ability against salinity-triggered oxidative stress by increasing the effectiveness of POD, SOD, and CAT enzymes in Triticum aestivum [252].

6. Hazardous Effects of N Fertilizers

Nutrient deficiency and/or availability are crucial to plant production [255,256]. One of the severe issues the world faces is the adverse effects of fertilizers on the environment. Farmers have been using fertilizers since ancient times, but long-term use has affected soil fertility adversely. Due to over-fertilization in many agricultural areas, environmental pollution was established [257]. N pollution in groundwater is a severe problem worldwide, affecting animal and human health; in addition, NO3 leaching decreases N availability for crops and can cause water pollution [258]. The improper application of fertilizers damages the farmlands, leading to heavy metal contamination and soil erosion. While ammonium sulfate causes soil pollution, nitrogen oxides and NH4+ cause air pollution, and NO3 is the final breakdown product accumulating in groundwater [259,260]. Since NO3 levels above 10 mg L−1 N cause oxygen debt in the blood, the U.S. Environmental Protection Agency set human drinking water to maintain nitrate levels below a maximum contaminant level (MCL) at a critical value of 10 mg L−1 N [261]. In Argentina, the national food law set the limit for human consumption at 11.3 mg L−1 N [258]. Consumption of NO3-rich water has been associated with methemoglobinemia (blue child syndrome) and different forms of cancer [262,263,264]. Furthermore, diabetes, adverse reproductive outcomes (especially neural tube defects), and thyroid conditions are related to NO3-contaminated drinking water.
A notable study by Maghanga et al. (2013) was conducted to monitor changes in surface water NO3 levels in ten rivers within a Kenyan tea plantation for three years. Water samples were obtained before and after fertilizer application in 2004, 2005, and 2006. For three years, there was no established trend between surface water nitrate levels and the time of fertilizer applications. However, the highest nitrate–nitrogen levels were in thr river Temochewa during the first fertilizer applications (4.9 mg L−1 to 8.2 mg L−1). Furthermore, fertilizer application increased NO3 levels, and the study indicated that initial nitrate–nitrogen levels in most rivers were high, causing surface water contamination [261].

7. Promoting N-Fixation through Diazotrophic Microbiota

Rapid human population growth and declining agricultural soils intensify resource competition, affecting human food hunger and sustainable agriculture systems worldwide [5]. Hence, developing alternative strategies to increase the current food demand is an inevitable need [265]. Accordingly, crop productivity is the pillar of nutritional food security and heavily depends on applying N fertilizers [266]. Applying N fertilizers consumes vast amounts of energy and causes excessive harm to the environment and health, expensive costs, soil fertility reduction, and other negative consequences [267,268]. Therefore, increasing plant NUE is essential for developing sustainable agriculture [269]. Ultimately, N fixation has multiple benefits through improving the yields of landraces that often receive little attention from breeders and empowering farmers using a sustainable and less expensive strategy. Additionally, it reduces the need for fossil fuels requiring synthetic N fertilizers [270,271]. Overall, the symbiotic relationship between leguminous plants and Rhizobia is a powerful adaptation strategy that enables these plants to reduce their dependence on synthetic fertilizers and thrive in N-deficient soils. Improving the microbial activities that boost plant growth is necessary to increase food production [272]. In that sense, several beneficial mutualistic microbes have been discovered; however, their dependable utilization as biofertilizers in soil conditions is still challenged. Thus, improving the microbial inoculants and exploring their diversity performance in the external habitat could lead to understanding the knowledge gap between the microbial performance in vitro and the external habitat [273].
Bacterial endophytes associated with crops, i.e., corn and wheat, can produce various sugar compositions, such as arabinose [274], to stimulate the production of a thick gelatinous layer (mucilage) around the root systems of these crops. This layer allows crops to better uptake fertilizer by forming a continuous boundary layer between roots and soil particles, protecting the root system from drought and supporting other microbes (Figure 5). Most interestingly, mucilage structures associated with complex microbiota contribute to nutrient acquisition and play a vital role in plant growth and defense [271].
Improving the sustainability of biological nitrogen fixation (BNF) to cereal crops is a future reason to enhance crop productivity, representing a significant breakthrough in N-fixation research. To achieve this goal, engineering key regulators of BNF by activating the symbiotic signaling pathway between diazotrophic bacteria and cereal crops is a promising strategy for sustainable development.
Fortunately, establishing new omics techniques such as metabolomics, metagenomics, metatranscriptomics, and metaproteomics will help in-depth research and identify the rhizospheric microbes using next-generation sequencing approaches [275]. Accordingly, manipulating the promising microbes that replace synthetic fertilizers is a solution. Rhizobium–plant symbiosis is now becoming challenging with diverse genetic modifiers of the symbiosis relationship not only in the genomic variability of partners but also in the soil microbiota. In that sense, there is a future direction to improve the specificity of the partnership during symbiotic colonization using different methods (e.g., axenic cultures) by eliminating other microbes. Therefore, identifying and developing bioinoculant strains and genotypes by adapting and evolving elite strains in both lab-scale analyses and trials to open-field applications to increase crop yield becomes crucial. Figure 5 shows how microbes can help plants increase their N availability by colonizing their roots, releasing exudates, and enhancing nutrient uptake.

8. Microbial Alternatives to Synthetic Fertilizers

Fertile soil is a primary component in the backbone of sustainable agriculture; however, land degradation and rapid desertification cause an estimated global loss of 24 billion tons of fertile arable land [276]. Noticeably, eutrophication and decline of soil fertility result from massive application of chemical fertilizers, causing significant environmental concern [277]. Moreover, with the limited N resources and increasing agricultural demands for N supplies, there is a need to discover diverse mutualistic interactions between plant roots and rhizosphere microbiome (Figure 5).
Soils are habitats for diverse microorganisms that are either harmful (pathogenic) or beneficial (symbiotic) to plants, comprising organic substances, water, and nutrients. Food production requires essential nutrients, metabolites, and water that the soil provides to plants [5,278,279]. Besides the nutrient resources in the soil, plants attract microbes via root exudates, root border cells, and mucilage formation that serve as food for the rhizosphere microbiome and their assembly [280]. Interestingly, soil contains an extensive reservoir of microorganisms (1 × 109 microbial cells g−1 dry soil) and microbial diversity (1 × 105 microbial species g−1 dry soil) [281,282]. Rhizospheric microbes are crucial in diverse biological processes, including soil structure, climate regulation, disease control, and organic matter decomposition. They also induce plant growth via nutrient uptake, cycling water availability, and the formation of stable aggregates to reduce the risks of soil erosion [283,284,285]. In addition, these microbes can improve plant health and serve as biocontrol agents against plant pathogens as well as biofertilizers (i.e., Rhizobium, Azotobacter, and Bacillus subtilis) [273,286]. Accordingly, elucidating the functional role of these microbial communities has garnered enormous efforts from botanical scientists over two decades to eliminate hunger and improve soil fertility by reducing synthetic fertilizers [287,288,289,290]. Fortunately, legume plants represent a solution, including fava beans, soybeans, lentils, and cowpea. They recruit symbiotic Rhizobia from the soil into nodules BNF. BNF is an eco-friendly and manageable strategy by harnessing nitrogen-fixing endophytic and free-living rhizobacteria to increase N levels in agricultural land by converting N2 to the fixed form, NH3 [291]. The biologically fixed amounts of atmospheric N into NH3 inside root nodules by legumes account for about 65% of N utilization in global agriculture via N-fixing Rhizobia [292]. Accordingly, this process allows legume plants to grow well in N-deficient soils, eliminating N fertilizer application [293]. The inoculation of Rhizobia strains (native or commercial) was studied on the growth and nodulation of three cowpea (Vigna unguiculata) genotypes in semiarid regions of Kenya. Field trials were performed in a randomized complete block design with three replications, and the symbiotic efficiency (SE) of Rhizobium isolates was evaluated. In the field, Rhizobia inoculation significantly increased nodulation and shoot DW compared to the uninoculated controls. Interestingly, Rhizobia inoculation significantly increased yields, whereas inoculation with native isolates recorded a 22.7% increase compared to uninoculated control in the first season and a 28.6% increase in yield in the second season. This study concluded that the efficient native Rhizobia in smallholder farms has promising potential to improve cowpea yield under a changing climate. In particular, R. tropici clone H53, Mesorhizobium sp. WSM3874, and R. pusense strain Nak353 showed more superiority in all the tested parameters [294]. A recent study revealed that plant exudates play a role in the development of bacterial biofilms (multicellular communities of microorganisms) in the soil, and creating microaerophilic conditions improved the nitrogen-fixing efficacy of diazotrophic bacteria in the soil [295] and exhibited enhanced flavonoid compounds. This study also hypothesized that manipulating flavonoid synthetic pathways leads to the induction of BNF in cereals through biofilm formation in soil diazotrophs [296,297]. N-limitation induces the production of exopolysaccharides (EPS), which may function as a barrier to block excess oxygen in the air and provide a suitable microaerobic condition for bacteria inside the EPS to fix N [298,299]. A large amount of EPS was reported in response to N limitation by cellulolytic bacteria (e.g., Bacillus, Pedobacter, Chryseobacterium, and Flavobacterium [300].
Applying the symbiotic association between Rhizobia and leguminous plants is a significant research area to reduce the dependency on chemical fertilizers. Bradyrhizobium, Azospirillum, and Rhizobium had high plant-growth-promoting capacity and N-fixing efficiency, indicating abundant functional microbial resources in extreme soil environments [301,302]. On the other hand, inoculation of Azospirillum, Burkholderia, and other N-fixing bacteria had little success in the field [303]. Thus, developing a compatible host for the indigenous N-fixing bacteria is critical for colonizing the host and improving nitrogenase activity.

9. Genomic Approaches for Improving N-Fixation

The identification of new genes involved in N fixation, investigation of the genome organization of N-fixing species, and the discovery of novel diazotrophic species have all been made possible by plant and microbial genomics. The primary genes responsible for N fixation encode nitrogenase enzymes, which play a crucial role in converting N2 into NH4+ [304,305]. These enzymes are classified into several types based on their structural, evolutionary, and functional characteristics. The first three forms of nitrogenase share similarities in their architecture, evolutionary origins, and mechanisms. However, they differ in the type of metal they use as a cofactor, either vanadium or molybdenum, or they may rely solely on iron as their cofactor. In contrast, the fourth type of nitrogenase operates differently; it utilizes a molybdopterin-containing system and depends on the presence of superoxide. This distinct form of nitrogenase was initially identified in a specific species called Streptomyces thermoautotrophicus. The complete genomic sequence of S. thermoautotrophicus and the precise amino acid sequence of this fourth type of nitrogenase have not been determined [306,307]. The fact that the inventory and distribution of species encoding for the fourth kind of nitrogenase are unknown highlights how few diazotrophs’ genomes have been sequenced. A. vinelandii is one of the extensively researched diazotroph with sequenced genomes that can fix N in an aerobic culture. It has served as a model for N-fixation for decades due to several factors, including (i) its genetic adaptability; (ii) its capacity to fix N in aerobic growth conditions; (iii) its nutritional flexibility, which is demonstrated by its capacity to fix N via three different pathways; and (iv) its capacity to adapt its metabolism to a variety of nutrients and media additives; and (v) the genome’s comprehensive sequencing and manual editing [308,309]. A. vinelandii has a single circular chromosome with 5,365,318 base pairs in its genome, which is expected to code for 5051 proteins [309]. The diazotroph Pseudomonas stutzeri is the closest companion of A. vinelandii with a sequenced genome [310]. Moreover, P. stutzeri grows aerobically and can only fix N in microaerobic conditions [311,312]. Alternatively, in ambient oxygen concentrations (20% O2), A. vinelandii can catalyze N fixation in an oxygen-sensitive process [313,314]. Its respiratory protection system makes it feasible to undertake two incompatible biological processes, oxidative phosphorylation and N fixation, concurrently. A. vinelandii may modify its respiration rate to maintain a low amount of cytoplasmic oxygen while N-fixing is taking place [315]. Five terminal oxidases were found in the genome, along with additional NADH oxidoreductases and other respiratory complexes that provide electrons to terminal oxidases, raising oxygen consumption [309]. Some of these genes, such as cydAB I, have been linked to respiratory protection and are crucial for aerobic N-fixation [306,307]. The discovery of novel N-fixing organisms and the quantity and kind of nitrogenases coding in these species may also be targeted by genome sequencing utilizing NifD sequences as queries [316,317,318]. The genomes of diazotrophs also encode for other proteins comparable to NifD in addition to the Nif/Vnf/AnfD sequences [318].

10. Conclusions and Future Perspectives

N is an essential macronutrient required for the growth and development of all living organisms. N uptake, transport, assimilation, and metabolism are finely tuned in plants. Additionally, N serves as a key mediator in plant–microbe interactions, fostering beneficial relationships through symbiotic partnerships that enhance nutrient availability. Moreover, N plays a pivotal role in a plant’s response to various environmental stresses, such as drought, salinity, and disease. Understanding N dynamics is critical for optimizing crop yields and sustainability, as it impacts the health and productivity of individual plants and the entire ecosystem and food security.
Consequently, harnessing N in agriculture addresses the challenges of feeding a growing global population while minimizing the environmental footprint of farming practices. In this sense, improving NUE by plants is a promising strategy to reduce the negative impact of using synthetic fertilizers. BNF is an eco-friendly and manageable strategy to increase N levels in agricultural land. Thus, improving BNF by genome editing could replace synthetic fertilizers to reduce dependency on chemical fertilizers. In addition, discovering new plant microbiomes and understanding their structure, abundance, and diversity will open a new avenue to better understanding the agricultural scenario to sustain global food security under limited N conditions. Despite understanding the role of N in plant biology, the precise molecular mechanisms governing N sensing in plants are still unclear, which could lead to the development of more efficient N-use strategies.
Furthermore, investigating the effects of climate change on N availability and its environmental impact becomes increasingly crucial. For example, arctic land areas are expected to warm, resulting in significant permafrost thawing [319,320] and significant alterations in ecological functioning [321]. With thawing, a massive pool of immobile C trapped in permafrost [322] becomes available for breakdown and remobilization, resulting in CO2 and methane (CH4) emissions. The gaseous C leak from thawing permafrost is being extensively researched in order to determine the extent of the permafrost–carbon feedback to the climate [323,324]. Permafrost soils, on the other hand, are huge N reservoirs, with an approximate amount of 67 billion tons of total N in the top 3 m [325]. Organically bound N is mineralized upon thawing, resulting in the release of NH4+ and NO3 and driving nitrification and denitrification, the two primary processes producing greenhouse gas in soils [326], and these microorganisms can convert to harmful NO or N2 gas [327]. In addition to methane and carbon dioxide fluxes, warming arctic soils release N, which can increase N2O emissions, another potent greenhouse gas. The significance and amount of N2O released from arctic soils are unknown. There is a crucial need for a combination of in situ field investigations, laboratory experiments, and cutting-edge metagenomic techniques to better understand the role of microorganisms in climate change and seasonal changes in bacteria.

Author Contributions

O.Z. contributed to the introduction, N-uptake and transport, stress response, conclusion and future prospect. O.A.H. contributed to the N-fixation and genomic approached for improving N-fixation. A.A. contributed to the role of N in response to stress. M.A. contributed to N metabolism. M.S.Y. contribute to the GS and GOGAT and A.F.R. contributed to the Introduction, N- transport. D.S. revised the manuscript and Z.-C.Y. contributed to the writing, editing and revision the original and final draft of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

Ze-Chun Yuan’s lab appreciates the support from Agriculture and Agri-Food Canada A-base project funds (J-002488 and J-003105), Mitacs fund, the Natural Sciences and Engineering Research Council of Canada (NSERC) Strategic Partnership Grant, as well as the support from Ontario Greenhouse and Vegetable Growers Association. Omar Zayed is supported by funds to Danelle Seymour from California Department of Food and Agriculture (CDFA # 21-0001-052-SF) and the U.S. Department of Agriculture National Institute of Food and Agriculture grant # 2022-70029-38471.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lal, R. Restoring Soil Quality to Mitigate Soil Degradation. Sustainability 2015, 7, 5875–5895. [Google Scholar] [CrossRef]
  2. Kopittke, P.M.; Menzies, N.W.; Wang, P.; McKenna, B.A.; Lombi, E. Soil and the intensification of agriculture for global food security. Environ. Int. 2019, 132, 105078. [Google Scholar] [CrossRef] [PubMed]
  3. Pereira, P.; Bogunovic, I.; Muñoz-Rojas, M.; Brevik, E.C. Soil ecosystem services, sustainability, valuation and management. Curr. Opin. Environ. Sci. Health 2018, 5, 7–13. [Google Scholar] [CrossRef]
  4. Pimentel, D.; Burgess, M. Soil Erosion Threatens Food Production. Agriculture 2013, 3, 443–463. [Google Scholar] [CrossRef]
  5. Díaz-Rodríguez, A.M.; Salcedo Gastelum, L.A.; Félix Pablos, C.M.; Parra-Cota, F.I.; Santoyo, G.; Puente, M.L.; Bhattacharya, D.; Mukherjee, J.; de los Santos-Villalobos, S. The Current and Future Role of Microbial Culture Collections in Food Security Worldwide. Front. Sustain. Food Syst. 2021, 4, 614739. [Google Scholar] [CrossRef]
  6. Arora, N.K.; Khare, E.; Maheshwari, D.K. Plant Growth Promoting Rhizobacteria: Constraints in Bioformulation, Commercialization, and Future Strategies. In Plant Growth and Health Promoting Bacteria; Maheshwari, D.K., Ed.; Springer: Berlin/Heidelberg, Germany, 2011. [Google Scholar]
  7. Hirel, B.; Tétu, T.; Lea, P.J.; Dubois, F. Improving Nitrogen Use Efficiency in Crops for Sustainable Agriculture. Sustainability 2011, 3, 1452–1485. [Google Scholar] [CrossRef]
  8. Akinnifesi, F.K.; Ajayi, O.C.; Sileshi, G.; Chirwa, P.W.; Chianu, J. Fertiliser trees for sustainable food security in the maize-based production systems of East and Southern Africa. A review. Agron. Sustain. Dev. 2010, 30, 615–629. [Google Scholar] [CrossRef]
  9. Spiertz, H. Avenues to meet food security. The role of agronomy on solving complexity in food production and resource use. Eur. J. Agron. 2012, 43, 1–8. [Google Scholar] [CrossRef]
  10. Azadi, H.; Ho, P. Genetically modified and organic crops in developing countries: A review of options for food security. Biotechnol. Adv. 2010, 28, 160–168. [Google Scholar] [CrossRef]
  11. Singh, J.S.; Gupta, V.K. Soil microbial biomass: A key soil driver in management of ecosystem functioning. Sci. Total Environ. 2018, 634, 497–500. [Google Scholar] [CrossRef]
  12. Shrivastav, P.; Prasad, M.; Singh, T.B.; Yadav, A.; Goyal, D.; Ali, A.; Dantu, P.K. Role of nutrients in plant growth and development. In Contaminants in Agriculture: Sources, Impacts and Management; Springer: Cham, Switzerland, 2020; pp. 43–59. [Google Scholar]
  13. Zhang, S.; Liu, Y.; Du, M.; Shou, G.; Wang, Z.; Xu, G. Nitrogen as a regulator for flowering time in plant. Plant Soil 2022, 480, 1–29. [Google Scholar] [CrossRef]
  14. Ye, J.Y.; Tian, W.H.; Jin, C.W. Nitrogen in plants: From nutrition to the modulation of abiotic stress adaptation. Stress Biol. 2022, 2, 4. [Google Scholar] [CrossRef] [PubMed]
  15. Hu, B.; Chu, C. Nitrogen–phosphorus interplay: Old story with molecular tale. New Phytol. 2020, 225, 1455–1460. [Google Scholar] [CrossRef] [PubMed]
  16. Xu, X.; Du, X.; Wang, F.; Sha, J.; Chen, Q.; Tian, G.; Zhu, Z.; Ge, S.; Jiang, Y. Effects of potassium levels on plant growth, accumulation and distribution of carbon, and nitrate metabolism in apple dwarf rootstock seedlings. Front. Plant Sci. 2020, 11, 904. [Google Scholar] [CrossRef] [PubMed]
  17. Li, H.; Yu, M.; Du, X.-Q.; Wang, Z.-F.; Wu, W.-H.; Quintero, F.J.; Jin, X.-H.; Li, H.-D.; Wang, Y. NRT1.5/NPF7.3 functions as a proton-coupled H+/K+ antiporter for K+ loading into the xylem in Arabidopsis. Plant Cell 2017, 29, 2016–2026. [Google Scholar] [CrossRef] [PubMed]
  18. Kosegarten, H.U.; Hoffmann, B.; Mengel, K. Apoplastic pH and Fe3+ reduction in intact sunflower leaves. Plant Physiol. 1999, 121, 1069–1079. [Google Scholar] [CrossRef] [PubMed]
  19. Zhao, T.; LING, H.Q. Effects of pH and nitrogen forms on expression profiles of genes involved in iron homeostasis in tomato. Plant Cell Environ. 2007, 30, 518–527. [Google Scholar] [CrossRef]
  20. Scherer, H.W. Impact of sulfur on N2 fixation of legumes. In Sulfur Assimilation and Abiotic Stress in Plants; Springer: Berlin/Heidelberg, Germany, 2008; pp. 43–54. [Google Scholar]
  21. Clarkson, D.T.; Saker, L.R.; Purves, J.V. Depression of nitrate and ammonium transport in barley plants with diminished sulphate status. Evidence of co-regulation of nitrogen and sulphate intake. J. Exp. Bot. 1989, 40, 953–963. [Google Scholar] [CrossRef]
  22. Rietra, R.P.; Heinen, M.; Dimkpa, C.O.; Bindraban, P.S. Effects of nutrient antagonism and synergism on yield and fertilizer use efficiency. Commun. Soil Sci. Plant Anal. 2017, 48, 1895–1920. [Google Scholar] [CrossRef]
  23. Saud, S.; Fahad, S.; Yajun, C.; Ihsan, M.Z.; Hammad, H.M.; Nasim, W.; Jr, A.; Arif, M.; Alharby, H. Effects of nitrogen supply on water stress and recovery mechanisms in Kentucky bluegrass plants. Front. Plant Sci. 2017, 8, 983. [Google Scholar] [CrossRef]
  24. Masclaux-Daubresse, C.; Daniel-Vedele, F.; Dechorgnat, J.; Chardon, F.; Gaufichon, L.; Suzuki, A. Nitrogen uptake, assimilation and remobilization in plants: Challenges for sustainable and productive agriculture. Ann. Bot. 2010, 105, 1141–1157. [Google Scholar] [CrossRef] [PubMed]
  25. Sandhu, N.; Sethi, M.; Kumar, A.; Dang, D.; Singh, J.; Chhuneja, P. Biochemical and genetic approaches improving nitrogen use efficiency in cereal crops: A review. Front. Plant Sci. 2021, 12, 657629. [Google Scholar] [CrossRef] [PubMed]
  26. Delin, S.; Stenberg, B.; Nyberg, A.; Brohede, L. Potential methods for estimating nitrogen fertilizer value of organic residues. Soil Use Manag. 2012, 28, 283–291. [Google Scholar] [CrossRef]
  27. Yao, X.; Nie, J.; Bai, R.; Sui, X. Amino acid transporters in plants: Identification and function. Plants 2020, 9, 972. [Google Scholar] [CrossRef] [PubMed]
  28. Li, F.; Dong, C.; Yang, T.; Bao, S.; Fang, W.; Lucas, W.J.; Zhang, Z. The tea plant CsLHT1 and CsLHT6 transporters take up amino acids, as a nitrogen source, from the soil of organic tea plantations. Hortic. Res. 2021, 8, 178. [Google Scholar] [CrossRef] [PubMed]
  29. Feng, H.; Fan, X.; Miller, A.J.; Xu, G. Plant nitrogen uptake and assimilation: Regulation of cellular pH homeostasis. J. Exp. Bot. 2020, 71, 4380–4392. [Google Scholar] [CrossRef] [PubMed]
  30. Forde, B.G. Nitrate transporters in plants: Structure, function and regulation. Biochim. Et Biophys. Acta (BBA)-Biomembr. 2000, 1465, 219–235. [Google Scholar] [CrossRef]
  31. Iqbal, A.; Qiang, D.; Zhun, W.; Xiangru, W.; Huiping, G.; Hengheng, Z.; Nianchang, P.; Xiling, Z.; Meizhen, S. Growth and nitrogen metabolism are associated with nitrogen-use efficiency in cotton genotypes. Plant Physiol. Biochem. 2020, 149, 61–74. [Google Scholar] [CrossRef]
  32. Léran, S.; Varala, K.; Boyer, J.-C.; Chiurazzi, M.; Crawford, N.; Daniel-Vedele, F.; David, L.; Dickstein, R.; Fernandez, E.; Forde, B. A unified nomenclature of nitrate transporter 1/peptide transporter family members in plants. Trends Plant Sci. 2014, 19, 5–9. [Google Scholar] [CrossRef]
  33. Wang, Y.-Y.; Hsu, P.-K.; Tsay, Y.-F. Uptake, allocation and signaling of nitrate. Trends Plant Sci. 2012, 17, 458–467. [Google Scholar] [CrossRef]
  34. Wang, Y.-Y.; Cheng, Y.-H.; Chen, K.-E.; Tsay, Y.-F. Nitrate transport, signaling, and use efficiency. Annu. Rev. Plant Biol. 2018, 69, 85–122. [Google Scholar] [CrossRef] [PubMed]
  35. Kiba, T.; Feria-Bourrellier, A.-B.; Lafouge, F.; Lezhneva, L.; Boutet-Mercey, S.; Orsel, M.; Bréhaut, V.; Miller, A.; Daniel-Vedele, F.; Sakakibara, H. The Arabidopsis nitrate transporter NRT2.4 plays a double role in roots and shoots of nitrogen-starved plants. Plant Cell 2012, 24, 245–258. [Google Scholar] [CrossRef] [PubMed]
  36. Chattha, M.S.; Ali, Q.; Haroon, M.; Afzal, M.J.; Javed, T.; Hussain, S.; Mahmood, T.; Solanki, M.K.; Umar, A.; Abbas, W. Enhancement of nitrogen use efficiency through agronomic and molecular based approaches in cotton. Front. Plant Sci. 2022, 13, 994306. [Google Scholar] [CrossRef] [PubMed]
  37. Bouguyon, E.; Brun, F.; Meynard, D.; Kubeš, M.; Pervent, M.; Leran, S.; Lacombe, B.; Krouk, G.; Guiderdoni, E.; Zažímalová, E. Multiple mechanisms of nitrate sensing by Arabidopsis nitrate transceptor NRT1. 1. Nat. Plants 2015, 1, 15015. [Google Scholar] [CrossRef] [PubMed]
  38. Ho, C.-H.; Lin, S.-H.; Hu, H.-C.; Tsay, Y.-F. CHL1 functions as a nitrate sensor in plants. Cell 2009, 138, 1184–1194. [Google Scholar] [CrossRef] [PubMed]
  39. Hu, H.C.; Wang, Y.Y.; Tsay, Y.F. AtCIPK8, a CBL-interacting protein kinase, regulates the low-affinity phase of the primary nitrate response. Plant J. 2009, 57, 264–278. [Google Scholar] [CrossRef] [PubMed]
  40. Maghiaoui, A.; Gojon, A.; Bach, L. NRT1. 1-centered nitrate signaling in plants. J. Exp. Bot. 2020, 71, 6226–6237. [Google Scholar] [CrossRef]
  41. Izmailov, S.; Nikitin, A. Nitrate signaling in plants: Mechanisms of implementation. Russ. J. Plant Physiol. 2020, 67, 31–44. [Google Scholar] [CrossRef]
  42. O’Brien, J.A.; Vega, A.; Bouguyon, E.; Krouk, G.; Gojon, A.; Coruzzi, G.; Gutiérrez, R.A. Nitrate transport, sensing, and responses in plants. Mol. Plant 2016, 9, 837–856. [Google Scholar] [CrossRef]
  43. Fang, X.Z.; Fang, S.Q.; Ye, Z.Q.; Liu, D.; Zhao, K.L.; Jin, C.W. NRT1.1 dual-affinity nitrate transport/signalling and its roles in plant abiotic stress resistance. Front. Plant Sci. 2021, 12, 715694. [Google Scholar] [CrossRef]
  44. Chen, C.-Z.; Lv, X.-F.; Li, J.-Y.; Yi, H.-Y.; Gong, J.-M. Arabidopsis NRT1.5 is another essential component in the regulation of nitrate reallocation and stress tolerance. Plant Physiol. 2012, 159, 1582–1590. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, G.-B.; Meng, S.; Gong, J.-M. The expected and unexpected roles of nitrate transporters in plant abiotic stress resistance and their regulation. Int. J. Mol. Sci. 2018, 19, 3535. [Google Scholar] [CrossRef] [PubMed]
  46. Nazish, T.; Arshad, M.; Jan, S.U.; Javaid, A.; Khan, M.H.; Naeem, M.A.; Baber, M.; Ali, M. Transporters and transcription factors gene families involved in improving nitrogen use efficiency (NUE) and assimilation in rice (Oryza sativa L.). Transgenic Res. 2021, 31, 23–42. [Google Scholar] [CrossRef] [PubMed]
  47. Guo, B.; Li, Y.; Wang, S.; Li, D.; Lv, C.; Xu, R. Characterization of the Nitrate Transporter gene family and functional identification of HvNRT2.1 in barley (Hordeum vulgare L.). PLoS ONE 2020, 15, e0232056. [Google Scholar] [CrossRef] [PubMed]
  48. Aluko, O.O.; Kant, S.; Adedire, O.M.; Li, C.; Yuan, G.; Liu, H.; Wang, Q. Unlocking the potentials of nitrate transporters at improving plant nitrogen use efficiency. Front. Plant Sci. 2023, 14, 1074839. [Google Scholar] [CrossRef]
  49. Ji, L.; Song, L.; Zou, L.; Li, S.; Zhang, R.; Yang, J.; Wang, C.; Zhang, Y.; Wang, X.; Yun, L. Cassava Nitrate Transporter NPF5.4 Promotes Both Yield Potential and Salt Tolerance in Rice. Mol. Breed. 2022. preprint. [Google Scholar] [CrossRef]
  50. Zhao, L.; Chen, P.; Liu, P.; Song, Y.; Zhang, D. Genetic effects and expression patterns of the nitrate transporter (NRT) gene family in Populus tomentosa. Front. Plant Sci. 2021, 12, 661635. [Google Scholar] [CrossRef] [PubMed]
  51. Qi, J.; Yu, L.; Ding, J.; Ji, C.; Wang, S.; Wang, C.; Ding, G.; Shi, L.; Xu, F.; Cai, H. Transcription factor OsSNAC1 positively regulates nitrate transporter gene expression in rice. Plant Physiol. 2023, 192, 2923–2942. [Google Scholar]
  52. Peña, P.A.; Quach, T.; Sato, S.; Ge, Z.; Nersesian, N.; Changa, T.; Dweikat, I.; Soundararajan, M.; Clemente, T.E. Expression of the maize Dof1 transcription factor in wheat and sorghum. Front. Plant Sci. 2017, 8, 434. [Google Scholar] [CrossRef]
  53. Cao, H.; Qi, S.; Sun, M.; Li, Z.; Yang, Y.; Crawford, N.M.; Wang, Y. Overexpression of the maize ZmNLP6 and ZmNLP8 can complement the Arabidopsis nitrate regulatory mutant nlp7 by restoring nitrate signaling and assimilation. Front. Plant Sci. 2017, 8, 1703. [Google Scholar] [CrossRef]
  54. Alvarez, J.M.; Riveras, E.; Vidal, E.A.; Gras, D.E.; Contreras-López, O.; Tamayo, K.P.; Aceituno, F.; Gómez, I.; Ruffel, S.; Lejay, L. Systems approach identifies TGA1 and TGA4 transcription factors as important regulatory components of the nitrate response of Arabidopsis thaliana roots. Plant J. 2014, 80, 1–13. [Google Scholar] [CrossRef] [PubMed]
  55. Liu, M.; Chang, W.; Fan, Y.; Sun, W.; Qu, C.; Zhang, K.; Liu, L.; Xu, X.; Tang, Z.; Li, J. Genome-wide identification and characterization of NODULE-INCEPTION-like protein (NLP) family genes in Brassica napus. Int. J. Mol. Sci. 2018, 19, 2270. [Google Scholar] [CrossRef] [PubMed]
  56. Chen, X.; Yao, Q.; Gao, X.; Jiang, C.; Harberd, N.P.; Fu, X. Shoot-to-root mobile transcription factor HY5 coordinates plant carbon and nitrogen acquisition. Curr. Biol. 2016, 26, 640–646. [Google Scholar] [CrossRef] [PubMed]
  57. Rubin, G.; Tohge, T.; Matsuda, F.; Saito, K.; Scheible, W.-R.d. Members of the LBD family of transcription factors repress anthocyanin synthesis and affect additional nitrogen responses in Arabidopsis. Plant Cell 2009, 21, 3567–3584. [Google Scholar] [CrossRef] [PubMed]
  58. Vidal, E.A.; Álvarez, J.M.; Gutiérrez, R.A. Nitrate regulation of AFB3 and NAC4 gene expression in Arabidopsis roots depends on NRT1. 1 nitrate transport function. Plant Signal. Behav. 2014, 9, e28501. [Google Scholar] [CrossRef] [PubMed]
  59. Du, X.-Q.; Wang, F.-L.; Li, H.; Jing, S.; Yu, M.; Li, J.; Wu, W.-H.; Kudla, J.; Wang, Y. The transcription factor MYB59 regulates K+/NO3− translocation in the Arabidopsis response to low K+ stress. Plant Cell 2019, 31, 699–714. [Google Scholar] [CrossRef] [PubMed]
  60. Medici, A.; Marshall-Colon, A.; Ronzier, E.; Szponarski, W.; Wang, R.; Gojon, A.; Crawford, N.M.; Ruffel, S.; Coruzzi, G.M.; Krouk, G. AtNIGT1/HRS1 integrates nitrate and phosphate signals at the Arabidopsis root tip. Nat. Commun. 2015, 6, 6274. [Google Scholar] [CrossRef]
  61. Guan, P.; Wang, R.; Nacry, P.; Breton, G.; Kay, S.A.; Pruneda-Paz, J.L.; Davani, A.; Crawford, N.M. Nitrate foraging by Arabidopsis roots is mediated by the transcription factor TCP20 through the systemic signaling pathway. Proc. Natl. Acad. Sci. USA 2014, 111, 15267–15272. [Google Scholar] [CrossRef]
  62. Dash, M.; Yordanov, Y.S.; Georgieva, T.; Kumari, S.; Wei, H.; Busov, V. A systems biology approach identifies new regulators of poplar root development under low nitrogen. Plant J. 2015, 84, 335–346. [Google Scholar] [CrossRef]
  63. Shin, J.M.; Chung, K.; Sakamoto, S.; Kojima, S.; Yeh, C.-M.; Ikeda, M.; Mitsuda, N.; Ohme-Takagi, M. The chimeric repressor for the GATA4 transcription factor improves tolerance to nitrogen deficiency in Arabidopsis. Plant Biotechnol. 2017, 34, 151–158. [Google Scholar] [CrossRef]
  64. Chen, M.; Zhu, K.; Xie, J.; Liu, J.; Tan, P.; Peng, F. Genome-Wide Identification and Expression Analysis of AMT and NRT Gene Family in Pecan (Carya illinoinensis) Seedlings Revealed a Preference for NH4+-N. Int. J. Mol. Sci. 2022, 23, 13314. [Google Scholar] [CrossRef] [PubMed]
  65. Schuldiner, S.; Shirvan, A.; Linial, M. Vesicular neurotransmitter transporters: From bacteria to humans. Physiol. Rev. 1995, 75, 369–392. [Google Scholar] [CrossRef] [PubMed]
  66. Tsay, Y.-F.; Chiu, C.-C.; Tsai, C.-B.; Ho, C.-H.; Hsu, P.-K. Nitrate transporters and peptide transporters. FEBS Lett. 2007, 581, 2290–2300. [Google Scholar] [CrossRef] [PubMed]
  67. Sun, J.; Bankston, J.R.; Payandeh, J.; Hinds, T.R.; Zagotta, W.N.; Zheng, N. Crystal structure of the plant dual-affinity nitrate transporter NRT1. 1. Nature 2014, 507, 73–77. [Google Scholar] [CrossRef] [PubMed]
  68. Madej, M.G.; Dang, S.; Yan, N.; Kaback, H.R. Evolutionary mix-and-match with MFS transporters. Proc. Natl. Acad. Sci. USA 2013, 110, 5870–5874. [Google Scholar] [CrossRef] [PubMed]
  69. Orsel, M.; Chopin, F.; Leleu, O.; Smith, S.J.; Krapp, A.; Daniel-Vedele, F.o.; Miller, A.J. Characterization of a Two-Component High-Affinity Nitrate Uptake System in Arabidopsis. Physiology and Protein-Protein Interaction. Plant Physiol. 2006, 142, 1304–1317. [Google Scholar] [CrossRef]
  70. Okamoto, M.; Kumar, A.; Li, W.; Wang, Y.; Siddiqi, M.Y.; Crawford, N.M.; Glass, A.D.M. High-Affinity Nitrate Transport in Roots of Arabidopsis Depends on Expression of the NAR2-Like Gene AtNRT3.1. Plant Physiol. 2006, 140, 1036–1046. [Google Scholar] [CrossRef] [PubMed]
  71. Benková, E.; Bielach, A. Lateral root organogenesis—From cell to organ. Curr. Opin. Plant Biol. 2010, 13, 677–683. [Google Scholar] [CrossRef]
  72. Yong, Z.; Kotur, Z.; Glass, A.D. Characterization of an intact two-component high-affinity nitrate transporter from Arabidopsis roots. Plant J. 2010, 63, 739–748. [Google Scholar] [CrossRef]
  73. Bowen, G.D.; Smith, S.E. The effects of mycorrhizas on nitrogen uptake by plants. Ecol. Bull. 1981, 33, 237–247. [Google Scholar]
  74. Öhlund, J.; Näsholm, T. Growth of conifer seedlings on organic and inorganic nitrogen sources. Tree Physiol. 2001, 21, 1319–1326. [Google Scholar] [CrossRef] [PubMed]
  75. Gazzarrini, S.; Lejay, L.; Gojon, A.; Ninnemann, O.; Frommer, W.B.; von Wirén, N. Three Functional Transporters for Constitutive, Diurnally Regulated, and Starvation-Induced Uptake of Ammonium into Arabidopsis Roots. Plant Cell 1999, 11, 937–947. [Google Scholar] [CrossRef] [PubMed]
  76. Bloom, A.J.; Meyerhoff, P.A.; Taylor, A.R.; Rost, T.L. Root Development and Absorption of Ammonium and Nitrate from the Rhizosphere. J. Plant Growth Regul. 2002, 21, 416–431. [Google Scholar] [CrossRef]
  77. Bittsánszky, A.; Pilinszky, K.; Gyulai, G.; Komives, T. Overcoming ammonium toxicity. Plant Sci. 2015, 231, 184–190. [Google Scholar] [CrossRef] [PubMed]
  78. Renard, J.J.; Calidonna, S.E.; Henley, M.V. Fate of ammonia in the atmosphere—A review for applicability to hazardous releases. J. Hazard. Mater. 2004, 108, 29–60. [Google Scholar] [CrossRef] [PubMed]
  79. Hachiya, T.; Watanabe, C.K.; Fujimoto, M.; Ishikawa, T.; Takahara, K.; Kawai-Yamada, M.; Uchimiya, H.; Uesono, Y.; Terashima, I.; Noguchi, K. Nitrate Addition Alleviates Ammonium Toxicity Without Lessening Ammonium Accumulation, Organic Acid Depletion and Inorganic Cation Depletion in Arabidopsis thaliana Shoots. Plant Cell Physiol. 2012, 53, 577–591. [Google Scholar] [CrossRef] [PubMed]
  80. von Wirén, N.; Gazzarrini, S.; Gojon, A.; Frommer, W.B. The molecular physiology of ammonium uptake and retrieval. Curr. Opin. Plant Biol. 2000, 3, 254–261. [Google Scholar] [CrossRef]
  81. Loqué, D.; von Wirén, N. Regulatory levels for the transport of ammonium in plant roots. J. Exp. Bot. 2004, 55, 1293–1305. [Google Scholar] [CrossRef]
  82. Glass, A.D.M.; Britto, D.T.; Kaiser, B.N.; Kinghorn, J.R.; Kronzucker, H.J.; Kumar, A.; Okamoto, M.; Rawat, S.; Siddiqi, M.Y.; Unkles, S.E.; et al. The regulation of nitrate and ammonium transport systems in plants. J. Exp. Bot. 2002, 53, 855–864. [Google Scholar] [CrossRef]
  83. Hao, D.-L.; Zhou, J.-Y.; Yang, S.-Y.; Qi, W.; Yang, K.-J.; Su, Y.-H. Function and Regulation of Ammonium Transporters in Plants. Int. J. Mol. Sci. 2020, 21, 3557. [Google Scholar] [CrossRef]
  84. Behie, S.W.; Bidochka, M.J. Ubiquity of Insect-Derived Nitrogen Transfer to Plants by Endophytic Insect-Pathogenic Fungi: An Additional Branch of the Soil Nitrogen Cycle. Appl. Environ. Microbiol. 2014, 80, 1553–1560. [Google Scholar] [CrossRef] [PubMed]
  85. Giehl, R.F.H.; Laginha, A.M.; Duan, F.; Rentsch, D.; Yuan, L.; von Wirén, N. A Critical Role of AMT2;1 in Root-To-Shoot Translocation of Ammonium in Arabidopsis. Mol. Plant 2017, 10, 1449–1460. [Google Scholar] [CrossRef] [PubMed]
  86. Couturier, J.; Montanini, B.; Martin, F.; Brun, A.; Blaudez, D.; Chalot, M. The expanded family of ammonium transporters in the perennial poplar plant. New Phytol. 2007, 174, 137–150. [Google Scholar] [CrossRef] [PubMed]
  87. Esmaeilzadeh-Salestani, K.; Samandari_Bahraseman, M.R.; Tohidfar, M.; Khaleghdoust, B.; Keres, I.; Mõttus, A.; Loit, E. Expression of AMT1;1 and AMT2;1 is stimulated by mineral nitrogen and reproductive growth stage in barley under field conditions. J. Plant Nutr. 2023, 46, 1246–1258. [Google Scholar] [CrossRef]
  88. Wu, X.; Yang, H.; Qu, C.; Xu, Z.; Li, W.; Hao, B.; Yang, C.; Sun, G.; Liu, G. Sequence and expression analysis of the AMT gene family in poplar. Front. Plant Sci. 2015, 6, 337. [Google Scholar] [CrossRef] [PubMed]
  89. Wu, X.X.; Yuan, D.P.; Chen, H.; Kumar, V.; Kang, S.M.; Jia, B.; Xuan, Y.H. Ammonium transporter 1 increases rice resistance to sheath blight by promoting nitrogen assimilation and ethylene signalling. Plant Biotechnol. J. 2022, 20, 1085–1097. [Google Scholar] [CrossRef]
  90. Pastor, V.; Gamir, J.; Camañes, G.; Cerezo, M.; Sánchez-Bel, P.; Flors, V. Disruption of the ammonium transporter AMT1.1 alters basal defenses generating resistance against Pseudomonas syringae and Plectosphaerella cucumerina. Front. Plant Sci. 2014, 5, 231. [Google Scholar] [CrossRef] [PubMed]
  91. Lima, J.E.; Kojima, S.; Takahashi, H.; von Wirén, N. Ammonium Triggers Lateral Root Branching in Arabidopsis in an AMMONIUM TRANSPORTER1;3-Dependent Manner. Plant Cell 2010, 22, 3621–3633. [Google Scholar] [CrossRef]
  92. Rogato, A.; D’Apuzzo, E.; Chiurazzi, M. The multiple plant response to high ammonium conditions. Plant Signal. Behav. 2010, 5, 1594–1596. [Google Scholar] [CrossRef]
  93. Rogato, A.; D’Apuzzo, E.; Barbulova, A.; Omrane, S.; Parlati, A.; Carfagna, S.; Costa, A.; Schiavo, F.L.; Esposito, S.; Chiurazzi, M. Characterization of a Developmental Root Response Caused by External Ammonium Supply in Lotus japonicus. Plant Physiol. 2010, 154, 784–795. [Google Scholar] [CrossRef]
  94. Cao, H.; Liu, Q.; Liu, X.; Ma, Z.; Zhang, J.; Li, X.; Shen, L.; Yuan, J.; Zhang, Q. Phosphatidic acid regulates ammonium uptake by interacting with AMMONIUM TRANSPORTER 1;1 in Arabidopsis. Plant Physiol. 2023. corrected proof. [Google Scholar] [CrossRef] [PubMed]
  95. Giovannetti, M.; Volpe, V.; Salvioli, A.; Bonfante, P. Chapter 7—Fungal and Plant Tools for the Uptake of Nutrients in Arbuscular Mycorrhizas: A Molecular View. In Mycorrhizal Mediation of Soil; Johnson, N.C., Gehring, C., Jansa, J., Eds.; Elsevier: Amsterdam, The Netherlands, 2017; pp. 107–128. [Google Scholar]
  96. Wu, Y.; Yang, W.; Wei, J.; Yoon, H.; An, G. Transcription Factor OsDOF18 Controls Ammonium Uptake by Inducing Ammonium Transporters in Rice Roots. Mol. Cells 2017, 40, 178–185. [Google Scholar] [PubMed]
  97. Yuan, L.; Gu, R.; Xuan, Y.; Smith-Valle, E.; Loqué, D.; Frommer, W.B.; von Wirén, N. Allosteric Regulation of Transport Activity by Heterotrimerization of Arabidopsis Ammonium Transporter Complexes in Vivo. Plant Cell 2013, 25, 974–984. [Google Scholar] [CrossRef] [PubMed]
  98. Lanquar, V.; Loqué, D.; Hörmann, F.; Yuan, L.; Bohner, A.; Engelsberger, W.R.; Lalonde, S.; Schulze, W.X.; von Wireén, N.; Frommer, W.B. Feedback Inhibition of Ammonium Uptake by a Phospho-Dependent Allosteric Mechanism in Arabidopsis. Plant Cell 2009, 21, 3610–3622. [Google Scholar] [CrossRef] [PubMed]
  99. Loqué, D.; Lalonde, S.; Looger, L.L.; von Wirén, N.; Frommer, W.B. A cytosolic trans-activation domain essential for ammonium uptake. Nature 2007, 446, 195–198. [Google Scholar] [CrossRef] [PubMed]
  100. Søgaard, R.; Alsterfjord, M.; MacAulay, N.; Zeuthen, T. Ammonium ion transport by the AMT/Rh homolog TaAMT1;1 is stimulated by acidic pH. Pflügers Arch.—Eur. J. Physiol. 2009, 458, 733–743. [Google Scholar] [CrossRef] [PubMed]
  101. Wang, Q.; Zhao, Y.; Luo, W.; Li, R.; He, Q.; Fang, X.; Michele, R.D.; Ast, C.; von Wirén, N.; Lin, J. Single-particle analysis reveals shutoff control of the Arabidopsis ammonium transporter AMT1;3 by clustering and internalization. Proc. Natl. Acad. Sci. USA 2013, 110, 13204–13209. [Google Scholar] [CrossRef]
  102. Koltun, A.; Maniero, R.A.; Vitti, M.; de Setta, N.; Giehl, R.F.H.; Lima, J.E.; Figueira, A. Functional characterization of the sugarcane (Saccharum spp.) ammonium transporter AMT2;1 suggests a role in ammonium root-to-shoot translocation. Front. Plant Sci. 2022, 13, 1039041. [Google Scholar] [CrossRef]
  103. Clegg, S.; Yu, F.; Griffiths, L.; Cole, J.A. The roles of the polytopic membrane proteins NarK, NarU and NirC in Escherichia coli K-12: Two nitrate and three nitrite transporters. Mol. Microbiol. 2002, 44, 143–155. [Google Scholar] [CrossRef]
  104. Wang, Y.; Li, W.; Siddiqi, Y.; Symington, V.F.; Kinghorn, J.R.; Unkles, S.E.; Glass, A.D. Nitrite transport is mediated by the nitrite-specific high-affinity NitA transporter and by nitrate transporters NrtA, NrtB in Aspergillus nidulans. Fungal Genet. Biol. 2008, 45, 94–102. [Google Scholar] [CrossRef]
  105. Kotur, Z.; Siddiqi, Y.M.; Glass, A.D. Characterization of nitrite uptake in Arabidopsis thaliana: Evidence for a nitrite-specific transporter. New Phytol. 2013, 200, 201–210. [Google Scholar] [CrossRef] [PubMed]
  106. Sugiura, M.; Georgescu, M.N.; Takahashi, M. A nitrite transporter associated with nitrite uptake by higher plant chloroplasts. Plant Cell Physiol. 2007, 48, 1022–1035. [Google Scholar] [CrossRef] [PubMed]
  107. Maeda, S.-I.; Konishi, M.; Yanagisawa, S.; Omata, T. Nitrite transport activity of a novel HPP family protein conserved in cyanobacteria and chloroplasts. Plant Cell Physiol. 2014, 55, 1311–1324. [Google Scholar] [CrossRef] [PubMed]
  108. Mistry, J.; Chuguransky, S.; Williams, L.; Qureshi, M.; Salazar, G.A.; Sonnhammer, E.L.; Tosatto, S.C.; Paladin, L.; Raj, S.; Richardson, L.J. Pfam: The protein families database in 2021. Nucleic Acids Res. 2021, 49, D412–D419. [Google Scholar] [CrossRef]
  109. Mengel, K.; Kirkby, E.A.; Kosegarten, H.; Appel, T. Nitrogen. In Principles of Plant Nutrition; Springer: Dordrecht, The Netherlands, 2001; pp. 397–434. [Google Scholar]
  110. Sparacino-Watkins, C.; Stolz, J.F.; Basu, P. Nitrate and periplasmic nitrate reductases. Chem. Soc. Rev. 2014, 43, 676–706. [Google Scholar] [CrossRef] [PubMed]
  111. Rosswall, T. Some Perspectives of the Major Biogeochemical Cycles; Likens, G.E., Ed.; John Wiley & Sons Ithaca: New York, NY, USA, 1981; Volume 2. [Google Scholar]
  112. Ward, M.H.; DeKok, T.M.; Levallois, P.; Brender, J.; Gulis, G.; Nolan, B.T.; VanDerslice, J. Workgroup report: Drinking-water nitrate and health—Recent findings and research needs. Environ. Health Perspect. 2005, 113, 1607–1614. [Google Scholar] [CrossRef] [PubMed]
  113. Powlson, D.S.; Addiscott, T.M.; Benjamin, N.; Cassman, K.G.; de Kok, T.M.; van Grinsven, H.; L’hirondel, J.L.; Avery, A.A.; Van Kessel, C. When does nitrate become a risk for humans? J. Environ. Qual. 2008, 37, 291–295. [Google Scholar] [CrossRef] [PubMed]
  114. Bonner, F.T.; Hughes, M.N. The aqueous solution chemistry of nitrogen in low positive oxidation states. Comments Inorg. Chem. 1988, 7, 215–234. [Google Scholar] [CrossRef]
  115. Jaffe, D.A. 12 The Nitrogen Cycle. In International Geophysics; Butcher, S.S., Charlson, R.J., Orians, G.H., Wolfe, G.V., Eds.; Academic Press: Cambridge, MA, USA, 1992; Volume 50, pp. 263–284. [Google Scholar]
  116. Canfield, D.E.; Glazer, A.N.; Falkowski, P.G. The evolution and future of Earth’s nitrogen cycle. Science 2010, 330, 192–196. [Google Scholar] [CrossRef]
  117. Ishii, S.; Ikeda, S.; Minamisawa, K.; Senoo, K. Nitrogen cycling in rice paddy environments: Past achievements and future challenges. Microbes Environ. 2011, 26, 282–292. [Google Scholar] [CrossRef]
  118. Elrys, A.S.; Zhu, Q.; Jiang, C.; Liu, J.; Sobhy, H.H.; Shen, Q.; Uwiragiye, Y.; Wu, Y.; El-Tarabily, K.A.; Meng, L. Global soil nitrogen cycle pattern and nitrogen enrichment effects: Tropical versus subtropical forests. Glob. Chang. Biol. 2023, 29, 1905–1921. [Google Scholar] [CrossRef] [PubMed]
  119. Joerger, R.; Jacobson, M.; Premakumar, R.; Wolfinger, E.; Bishop, P. Nucleotide sequence and mutational analysis of the structural genes (anfHDGK) for the second alternative nitrogenase from Azotobacter vinelandii. J. Bacteriol. 1989, 171, 1075–1086. [Google Scholar] [CrossRef] [PubMed]
  120. Joerger, R.; Loveless, T.; Pau, R.; Mitchenall, L.; Simon, B.; Bishop, P. Nucleotide sequences and mutational analysis of the structural genes for nitrogenase 2 of Azotobacter vinelandii. J. Bacteriol. 1990, 172, 3400–3408. [Google Scholar] [CrossRef] [PubMed]
  121. Fani, R.; Gallo, R.; Liò, P. Molecular evolution of nitrogen fixation: The evolutionary history of the nifD, nifK, nifE, and nifN genes. J. Mol. Evol. 2000, 51, 1–11. [Google Scholar] [CrossRef] [PubMed]
  122. Cantera, J.J.L.; Stein, L.Y. Molecular diversity of nitrite reductase genes (nirK) in nitrifying bacteria. Environ. Microbiol. 2007, 9, 765–776. [Google Scholar] [CrossRef] [PubMed]
  123. Kartal, B.; Maalcke, W.J.; De Almeida, N.M.; Cirpus, I.; Gloerich, J.; Geerts, W.; Op den Camp, H.J.; Harhangi, H.R.; Janssen-Megens, E.M.; Francoijs, K.-J. Molecular mechanism of anaerobic ammonium oxidation. Nature 2011, 479, 127–130. [Google Scholar] [CrossRef] [PubMed]
  124. Jepson, B.J.; Mohan, S.; Clarke, T.A.; Gates, A.J.; Cole, J.A.; Butler, C.S.; Butt, J.N.; Hemmings, A.M.; Richardson, D.J. Spectropotentiometric and structural analysis of the periplasmic nitrate reductase from Escherichia coli. J. Biol. Chem. 2007, 282, 6425–6437. [Google Scholar] [CrossRef]
  125. Heiss, B.; Frunzke, K.; Zumft, W. Formation of the NN bond from nitric oxide by a membrane-bound cytochrome bc complex of nitrate-respiring (denitrifying) Pseudomonas stutzeri. J. Bacteriol. 1989, 171, 3288–3297. [Google Scholar] [CrossRef]
  126. Suzuki, M.; Arai, H.; Ishii, M.; Igarashi, Y. Gene structure and expression profile of cytochrome bc nitric oxide reductase from Hydrogenobacter thermophilus TK-6. Biosci. Biotechnol. Biochem. 2006, 70, 1666–1671. [Google Scholar] [CrossRef]
  127. Ogawa, K.-i.; Akagawa, E.; Yamane, K.; Sun, Z.-W.; LaCelle, M.; Zuber, P.; Nakano, M.M. The nasB operon and nasA gene are required for nitrate/nitrite assimilation in Bacillus subtilis. J. Bacteriol. 1995, 177, 1409–1413. [Google Scholar] [CrossRef]
  128. Exley, G.E.; Colandene, J.D.; Garrett, R.H. Molecular cloning, characterization, and nucleotide sequence of nit-6, the structural gene for nitrite reductase in Neurospora crassa. J. Bacteriol. 1993, 175, 2379–2392. [Google Scholar] [CrossRef]
  129. Sohaskey, C.D.; Wayne, L.G. Role of narK2X and narGHJI inHypoxic Upregulation of Nitrate Reduction by Mycobacteriumtuberculosis. J. Bacteriol. 2003, 185, 7247–7256. [Google Scholar] [CrossRef] [PubMed]
  130. Stolz, J.F.; Basu, P. Evolution of nitrate reductase: Molecular and structural variations on a common function. Chembiochem 2002, 3, 198–206. [Google Scholar] [CrossRef] [PubMed]
  131. Kizawa, H.; Tomura, D.; Oda, M.; Fukamizu, A.; Hoshino, T.; Gotoh, O.; Yasui, T.; Shoun, H. Nucleotide sequence of the unique nitrate/nitrite-inducible cytochrome P-450 cDNA from Fusarium oxysporum. J. Biol. Chem. 1991, 266, 10632–10637. [Google Scholar] [CrossRef] [PubMed]
  132. Zhang, L.; Shoun, H. Purification and functional analysis of fungal nitric oxide reductase cytochrome P450nor. Methods Enzymol. 2008, 437, 117–133. [Google Scholar] [PubMed]
  133. COYLE, C.L.; ZUMFT, W.G.; KRONECK, P.M.; KÖRNER, H.; JAKOB, W. Nitrous oxide reductase from denitrifying: Pseudomonas perfectomarina purification and properties of a novel multicopper enzyme. Eur. J. Biochem. 1985, 153, 459–467. [Google Scholar] [CrossRef]
  134. Arai, H.; Mizutani, M.; Igarashi, Y. Transcriptional regulation of the nos genes for nitrous oxide reductase in Pseudomonas aeruginosa. Microbiology 2003, 149, 29–36. [Google Scholar] [CrossRef]
  135. Zumft, W.G.; Kroneck, P.M. Respiratory transformation of nitrous oxide (N2O) to dinitrogen by Bacteria and Archaea. Adv. Microb. Physiol. 2006, 52, 107–227. [Google Scholar]
  136. Keable, S.M.; Vertemara, J.; Zadvornyy, O.A.; Eilers, B.J.; Danyal, K.; Rasmussen, A.J.; De Gioia, L.; Zampella, G.; Seefeldt, L.C.; Peters, J.W. Structural characterization of the nitrogenase molybdenum-iron protein with the substrate acetylene trapped near the active site. J. Inorg. Biochem. 2018, 180, 129–134. [Google Scholar] [CrossRef]
  137. Hu, Y.; Lee, C.C.; Ribbe, M.W. Extending the carbon chain: Hydrocarbon formation catalyzed by vanadium/molybdenum nitrogenases. Science 2011, 333, 753–755. [Google Scholar] [CrossRef]
  138. Deistung, J.; Cannon, F.C.; Cannon, M.C.; Hill, S.; Thorneley, R. Electron transfer to nitrogenase in Klebsiella pneumoniae. nifF gene cloned and the gene product, a flavodoxin, purified. Biochem. J. 1985, 231, 743. [Google Scholar] [CrossRef] [PubMed]
  139. Fraser, J.A.; Davis, M.A.; Hynes, M.J. The formamidase gene of Aspergillus nidulans: Regulation by nitrogen metabolite repression and transcriptional interference by an overlapping upstream gene. Genetics 2001, 157, 119–131. [Google Scholar] [CrossRef] [PubMed]
  140. Maalcke, W.J.; Reimann, J.; De Vries, S.; Butt, J.N.; Dietl, A.; Kip, N.; Mersdorf, U.; Barends, T.R.; Jetten, M.S.; Keltjens, J.T. Characterization of anammox hydrazine dehydrogenase, a key N2-producing enzyme in the global nitrogen cycle. J. Biol. Chem. 2016, 291, 17077–17092. [Google Scholar] [CrossRef] [PubMed]
  141. Zhao, R.; Biddle, J.F.; Jørgensen, S.L. Introducing Candidatus Bathyanammoxibiaceae, a family of bacteria with the anammox potential present in both marine and terrestrial environments. ISME Commun. 2022, 2, 42. [Google Scholar] [CrossRef]
  142. Breton, S. The cellular physiology of carbonic anhydrases. Jop 2001, 2, 159–164. [Google Scholar] [PubMed]
  143. Pastorekova, S.; Parkkila, S.; Pastorek, J.; Supuran, C. Carbonic anhydrases: Current state of the art, therapeutic applications and future prospects. J. Enzym. Inhib. Med. Chem. 2004, 19, 199–229. [Google Scholar] [CrossRef] [PubMed]
  144. Harano, Y.; Suzuki, I.; Maeda, S.-i.; Kaneko, T.; Tabata, S.; Omata, T. Identification and nitrogen regulation of the cyanase gene from the cyanobacteria Synechocystis sp. strain PCC 6803 and Synechococcus sp. strain PCC 7942. J. Bacteriol. 1997, 179, 5744–5750. [Google Scholar] [CrossRef] [PubMed]
  145. Cantera, J.J.L.; Stein, L.Y. Role of nitrite reductase in the ammonia-oxidizing pathway of Nitrosomonas europaea. Arch. Microbiol. 2007, 188, 349–354. [Google Scholar] [CrossRef]
  146. Noji, S.; Nohno, T.; Saito, T.; Taniguchi, S. The narK gene product participates in nitrate transport induced in Escherichia coli nitrate-respiring cells. FEBS Lett. 1989, 252, 139–143. [Google Scholar] [CrossRef]
  147. Maeda, S.-I.; Omata, T. Nitrite transport activity of the ABC-type cyanate transporter of the cyanobacterium Synechococcus elongatus. J. Bacteriol. 2009, 191, 3265–3272. [Google Scholar] [CrossRef]
  148. Blasco, F.; Iobbi, C.; Ratouchniak, J.; Bonnefoy, V.; Chippaux, M. Nitrate reductases of Escherichia coli: Sequence of the second nitrate reductase and comparison with that encoded by the narGHJI operon. Mol. Gen. Genet. MGG 1990, 222, 104–111. [Google Scholar] [CrossRef] [PubMed]
  149. Okamoto, P.M.; Marzluf, G.A. Nitrate reductase of Neurospora crassa: The functional role of individual amino acids in the heme domain as examined by site-directed mutagenesis. Mol. Gen. Genet. MGG 1993, 240, 221–230. [Google Scholar] [CrossRef] [PubMed]
  150. Rubio, L.M.; Herrero, A.; Flores, E. A cyanobacterial narB gene encodes a ferredoxin-dependent nitrate reductase. Plant Mol. Biol. 1996, 30, 845–850. [Google Scholar] [CrossRef] [PubMed]
  151. Takahashi, M.; Sasaki, Y.; Ida, S.; Morikawa, H. Nitrite reductase gene enrichment improves assimilation of NO2 in Arabidopsis. Plant Physiol. 2001, 126, 731–741. [Google Scholar] [CrossRef] [PubMed]
  152. Harborne, N.R.; Griffiths, L.; Busby, S.J.; Cole, J.A. Transcriptional control, translation and function of the products of the five open reading frames of the Escherichia coli nir operon. Mol. Microbiol. 1992, 6, 2805–2813. [Google Scholar] [CrossRef] [PubMed]
  153. Einsle, O.; Messerschmidt, A.; Stach, P.; Bourenkov, G.P.; Bartunik, H.D.; Huber, R.; Kroneck, P.M. Structure of cytochrome c nitrite reductase. Nature 1999, 400, 476–480. [Google Scholar] [CrossRef] [PubMed]
  154. Gadda, G.; Francis, K. Nitronate monooxygenase, a model for anionic flavin semiquinone intermediates in oxidative catalysis. Arch. Biochem. Biophys. 2010, 493, 53–61. [Google Scholar] [CrossRef]
  155. Daubner, S.C.; Gadda, G.; Valley, M.P.; Fitzpatrick, P.F. Cloning of nitroalkane oxidase from Fusarium oxysporum identifies a new member of the acyl-CoA dehydrogenase superfamily. Proc. Natl. Acad. Sci. USA 2002, 99, 2702–2707. [Google Scholar] [CrossRef]
  156. Wolfe, M.T.; Heo, J.; Garavelli, J.S.; Ludden, P.W. Hydroxylamine Reductase Activity of the Hybrid Cluster Protein from Escherichia coli. J. Bacteriol. 2002, 184, 5898–5902. [Google Scholar] [CrossRef]
  157. Hommes, N.G.; Sayavedra-Soto, L.A.; Arp, D.J. Transcript Analysis of Multiple Copies of amo (Encoding Ammonia Monooxygenase) and hao (Encoding Hydroxylamine Oxidoreductase) in Nitrosomonas europaea. J. Bacteriol. 2001, 183, 1096–1100. [Google Scholar] [CrossRef]
  158. MOIR, J.W.B.; WEHRFRITZ, J.-M.; SPIRO, S.; RICHARDSON, D.J. The biochemical characterization of a novel non-haem-iron hydroxylamine oxidase from Paracoccus denitrificans GB17. Biochem. J. 1996, 319, 823–827. [Google Scholar] [CrossRef] [PubMed]
  159. Klotz, M.G.; Stein, L.Y. Genomics of Ammonia-Oxidizing Bacteria and Insights into Their Evolution. In Nitrification; ASM Press: Washington, DC, USA, 2011; pp. 57–94. [Google Scholar] [CrossRef]
  160. Durbecq, V.; Legrain, C.; Roovers, M.; Piérard, A.; Glansdorff, N. The carbamate kinase-like carbamoyl phosphate synthetase of the hyperthermophilic archaeon Pyrococcus furiosus, a missing link in the evolution of carbamoyl phosphate biosynthesis. Proc. Natl. Acad. Sci. USA 1997, 94, 12803–12808. [Google Scholar] [CrossRef] [PubMed]
  161. Ahuja, V.; Powers-Lee, S.G. Human carbamoyl-phosphate synthetase: Insight into N-acetylglutamate interaction and the functional effects of a common single nucleotide polymorphism. J. Inherit. Metab. Dis. 2008, 31, 481–491. [Google Scholar] [CrossRef] [PubMed]
  162. Belitsky, B.R.; Sonenshein, A.L. Role and Regulation of Bacillus subtilis Glutamate Dehydrogenase Genes. J. Bacteriol. 1998, 180, 6298–6305. [Google Scholar] [CrossRef] [PubMed]
  163. Frigerio, F.; Casimir, M.; Carobbio, S.; Maechler, P. Tissue specificity of mitochondrial glutamate pathways and the control of metabolic homeostasis. Biochim. Biophys. Acta (BBA)-Bioenerg. 2008, 1777, 965–972. [Google Scholar] [CrossRef] [PubMed]
  164. Riba, L.; Becerril, B.; Servín-González, L.; Valle, F.; Bolivar, F. Identification of a functional promoter for the Escherichia coli gdhA gene and its regulation. Gene 1988, 71, 233–246. [Google Scholar] [CrossRef] [PubMed]
  165. Newsholme, P.; Procopio, J.; Lima, M.M.R.; Pithon-Curi, T.C.; Curi, R. Glutamine and glutamate—their central role in cell metabolism and function. Cell Biochem. Funct. 2003, 21, 1–9. [Google Scholar] [CrossRef] [PubMed]
  166. Zhu, D.; Mukherjee, C.; Yang, Y.; Rios, B.E.; Gallagher, D.T.; Smith, N.N.; Biehl, E.R.; Hua, L. A new nitrilase from Bradyrhizobium japonicum USDA 110: Gene cloning, biochemical characterization and substrate specificity. J. Biotechnol. 2008, 133, 327–333. [Google Scholar] [CrossRef]
  167. Sonawane, A. Regulation of Pseudomonas putida Genes Involved in the Metabolism of Acidic Amino Acids. Ph.D. Thesis, Philipps-Universität Marburg, Marburg, Germany, 2003. [Google Scholar]
  168. Lancien, M.; Martin, M.; Hsieh, M.-H.; Leustek, T.; Goodman, H.; Coruzzi, G.M. Arabidopsis glt1-T mutant defines a role for NADH-GOGAT in the non-photorespiratory ammonium assimilatory pathway. Plant J. 2002, 29, 347–358. [Google Scholar] [CrossRef]
  169. Coschigano, K.T.; Melo-Oliveira, R.; Lim, J.; Coruzzi, G.M. Arabidopsis gls Mutants and Distinct Fd-GOGAT Genes: Implications for Photorespiration and Primary Nitrogen Assimilation. Plant Cell 1998, 10, 741–752. [Google Scholar] [CrossRef]
  170. Rees, E.; Siddiqui, R.A.; Köster, F.; Schneider, B.; Friedrich, B. Structural gene (nirS) for the cytochrome cd1 nitrite reductase of Alcaligenes eutrophus H16. Appl. Environ. Microbiol. 1997, 63, 800–802. [Google Scholar] [CrossRef] [PubMed]
  171. Hole, U.H.; Vollack, K.-U.; Zumft, W.G.; Eisenmann, E.; Siddiqui, R.A.; Friedrich, B.; Kroneck, P.M.H. Characterization of the membranous denitrification enzymes nitrite reductase (cytochrome cd1) and copper-containing nitrous oxide reductase from Thiobacillus denitrificans. Arch. Microbiol. 1996, 165, 55–61. [Google Scholar] [CrossRef] [PubMed]
  172. Dubey, R.S.; Srivastava, R.K.; Pessarakli, M. Physiological mechanisms of nitrogen absorption and assimilation in plants under stressful conditions. In Handbook of Plant and Crop Physiology; CRC Press: Boca Raton, FL, USA, 2021; pp. 579–616. [Google Scholar]
  173. Straub, T.; Ludewig, U.; Neuhäuser, B. The kinase CIPK23 inhibits ammonium transport in Arabidopsis thaliana. Plant Cell 2017, 29, 409–422. [Google Scholar] [CrossRef] [PubMed]
  174. Engels, C.; Marschner, H. Plant uptake and utilization of nitrogen. In Nitrogen Fertilization in the Environment; Marcel Dekker Inc.: New York, NY, USA, 1995; pp. 41–81. [Google Scholar]
  175. Garg, S. Role and hormonal regulation of nitrate reductase activity in higher plants: A review. Plant Sci. Feed 2013, 3, 13–20. [Google Scholar]
  176. Kaiser, W.M.; Huber, S.C. Post-translational regulation of nitrate reductase: Mechanism, physiological relevance and environmental triggers. J. Exp. Bot. 2001, 52, 1981–1989. [Google Scholar] [CrossRef] [PubMed]
  177. Nemie-Feyissa, D.; Królicka, A.; Førland, N.; Hansen, M.; Heidari, B.; Lillo, C. Post-translational control of nitrate reductase activity responding to light and photosynthesis evolved already in the early vascular plants. J. Plant Physiol. 2013, 170, 662–667. [Google Scholar] [CrossRef] [PubMed]
  178. Balotf, S.; Kavoosi, G.; Kholdebarin, B. Nitrate reductase, nitrite reductase, glutamine synthetase, and glutamate synthase expression and activity in response to different nitrogen sources in nitrogen-starved wheat seedlings. Biotechnol. Appl. Biochem. 2016, 63, 220–229. [Google Scholar] [CrossRef]
  179. Vega-Mas, I.; Cukier, C.; Coleto, I.; González-Murua, C.; Limami, A.M.; González-Moro, M.B.; Marino, D. Isotopic labelling reveals the efficient adaptation of wheat root TCA cycle flux modes to match carbon demand under ammonium nutrition. Sci. Rep. 2019, 9, 8925. [Google Scholar] [CrossRef]
  180. Fujita, T.; Beier, M.P.; Tabuchi-Kobayashi, M.; Hayatsu, Y.; Nakamura, H.; Umetsu-Ohashi, T.; Sasaki, K.; Ishiyama, K.; Murozuka, E.; Kojima, M. Cytosolic glutamine synthetase GS1; 3 is involved in rice grain ripening and germination. Front. Plant Sci. 2022, 13, 835835. [Google Scholar] [CrossRef]
  181. Lea, P.J.; Miflin, B.J. Nitrogen assimilation and its relevance to crop improvement. Annu. Plant Rev. 2011, 42, 1–40. [Google Scholar]
  182. Miflin, B.J.; Lea, P.J. The pathway of nitrogen assimilation in plants. Phytochemistry 1976, 15, 873–885. [Google Scholar] [CrossRef]
  183. Oaks, A.; Stulen, I.; Jones, K.; Winspear, M.J.; Misra, S.; Boesel, I.L. Enzymes of nitrogen assimilation in maize roots. Planta 1980, 148, 477–484. [Google Scholar] [CrossRef] [PubMed]
  184. Tamura, W.; Hidaka, Y.; Tabuchi, M.; Kojima, S.; Hayakawa, T.; Sato, T.; Obara, M.; Kojima, M.; Sakakibara, H.; Yamaya, T. Reverse genetics approach to characterize a function of NADH-glutamate synthase1 in rice plants. Amino Acids 2010, 39, 1003–1012. [Google Scholar] [CrossRef] [PubMed]
  185. Kojima, S.; Minagawa, H.; Yoshida, C.; Inoue, E.; Takahashi, H.; Ishiyama, K. Coregulation of glutamine synthetase1; 2 (GLN1; 2) and NADH-dependent glutamate synthase (GLT1) gene expression in Arabidopsis roots in response to ammonium supply. Front. Plant Sci. 2023, 14, 1127006. [Google Scholar] [CrossRef] [PubMed]
  186. Yasuda, T.; Konishi, N.; Kojima, S. Ammonium uptake capacity and response of cytosolic glutamine synthetase 1;2 to ammonium supply are key factors for the adaptation of ammonium nutrition in Arabidopsis thaliana. Soil Sci. Plant Nutr. 2017, 63, 553–560. [Google Scholar] [CrossRef]
  187. Konishi, N.; Saito, M.; Imagawa, F.; Kanno, K.; Yamaya, T.; Kojima, S. Cytosolic glutamine synthetase isozymes play redundant roles in ammonium assimilation under low-ammonium conditions in roots of Arabidopsis thaliana. Plant Cell Physiol. 2018, 59, 601–613. [Google Scholar] [CrossRef] [PubMed]
  188. Yamaya, T.; Kusano, M. Evidence supporting distinct functions of three cytosolic glutamine synthetases and two NADH-glutamate synthases in rice. J. Exp. Bot. 2014, 65, 5519–5525. [Google Scholar] [CrossRef]
  189. Kusano, M.; Fukushima, A.; Redestig, H.; Saito, K. Metabolomic approaches toward understanding nitrogen metabolism in plants. J. Exp. Bot. 2011, 62, 1439–1453. [Google Scholar] [CrossRef]
  190. Forde, B.G.; Lea, P.J. Glutamate in plants: Metabolism, regulation, and signalling. J. Exp. Bot. 2007, 58, 2339–2358. [Google Scholar] [CrossRef]
  191. Bowsher, C.G.; Lacey, A.E.; Hanke, G.T.; Clarkson, D.T.; Saker, L.R.; Stulen, I.; Emes, M.J. The effect of Glc6P uptake and its subsequent oxidation within pea root plastids on nitrite reduction and glutamate synthesis. J. Exp. Bot. 2007, 58, 1109–1118. [Google Scholar] [CrossRef]
  192. Habash, D.Z.; Bernard, S.; Schondelmaier, J.; Weyen, J.; Quarrie, S.A. The genetics of nitrogen use in hexaploid wheat: N utilisation, development and yield. Theor. Appl. Genet. 2007, 114, 403–419. [Google Scholar] [CrossRef] [PubMed]
  193. Yin, H.; Yang, F.; He, X.; Du, X.; Mu, P.; Ma, W. Advances in the functional study of glutamine synthetase in plant abiotic stress tolerance response. Crop J. 2022, 10, 917–923. [Google Scholar] [CrossRef]
  194. Prinsi, B.; Espen, L. Mineral nitrogen sources differently affect root glutamine synthetase isoforms and amino acid balance among organs in maize. BMC Plant Biol. 2015, 15, 96. [Google Scholar] [CrossRef] [PubMed]
  195. Zhang, Z.; Xiong, S.; Wei, Y.; Meng, X.; Wang, X.; Ma, X. The role of glutamine synthetase isozymes in enhancing nitrogen use efficiency of N-efficient winter wheat. Sci. Rep. 2017, 7, 1000. [Google Scholar] [CrossRef] [PubMed]
  196. Habash, D.; Massiah, A.; Rong, H.; Wallsgrove, R.; Leigh, R. The role of cytosolic glutamine synthetase in wheat. Ann. Appl. Biol. 2001, 138, 83–89. [Google Scholar] [CrossRef]
  197. Gadaleta, A.; Nigro, D.; Marcotuli, I.; Giancaspro, A.; Giove, S.L.; Blanco, A. Isolation and characterisation of cytosolic glutamine synthetase (GSe) genes and association with grain protein content in durum wheat. Crop Pasture Sci. 2014, 65, 38–45. [Google Scholar] [CrossRef]
  198. Tabuchi, M.; Abiko, T.; Yamaya, T. Assimilation of ammonium ions and reutilization of nitrogen in rice (Oryza sativa L.). J. Exp. Bot. 2007, 58, 2319–2327. [Google Scholar] [CrossRef]
  199. Ishiyama, K.; Inoue, E.; Watanabe-Takahashi, A.; Obara, M.; Yamaya, T.; Takahashi, H. Kinetic Properties and Ammonium-dependent Regulation of Cytosolic Isoenzymes of Glutamine Synthetase in Arabidopsis. J. Biol. Chem. 2004, 279, 16598–16605. [Google Scholar] [CrossRef]
  200. Yaakob, M.A.; Mohamed, R.M.S.R.; Al-Gheethi, A.; Aswathnarayana Gokare, R.; Ambati, R.R. Influence of nitrogen and phosphorus on microalgal growth, biomass, lipid, and fatty acid production: An overview. Cells 2021, 10, 393. [Google Scholar] [CrossRef]
  201. Morales-Sánchez, D.; Martinez-Rodriguez, O.A.; Kyndt, J.; Martinez, A. Heterotrophic growth of microalgae: Metabolic aspects. World J. Microbiol. Biotechnol. 2015, 31, 1–9. [Google Scholar] [CrossRef]
  202. Kumar, A.; Bera, S. Revisiting nitrogen utilization in algae: A review on the process of regulation and assimilation. Bioresour. Technol. Rep. 2020, 12, 100584. [Google Scholar] [CrossRef]
  203. Vega, J.M. Nitrogen and sulfur metabolism in microalgae and plants: 50 years of research. Prog. Bot. 2020, 81, 1–40. [Google Scholar]
  204. Zhang, H.; Zhu, J.; Gong, Z.; Zhu, J.-K. Abiotic stress responses in plants. Nat. Rev. Genet. 2022, 23, 104–119. [Google Scholar] [CrossRef] [PubMed]
  205. Sedri, M.H.; Amini, A.; Golchin, A. Evaluation of Nitrogen Effects on Yield and Drought Tolerance of Rainfed Wheat using Drought Stress Indices. J. Crop Sci. Biotechnol. 2019, 22, 235–242. [Google Scholar] [CrossRef]
  206. Song, J.; Wang, Y.; Pan, Y.; Pang, J.; Zhang, X.; Fan, J.; Zhang, Y. The influence of nitrogen availability on anatomical and physiological responses of Populus alba × P. glandulosa to drought stress. BMC Plant Biol. 2019, 19, 63. [Google Scholar] [CrossRef] [PubMed]
  207. Xu, N.; Guo, W.; Liu, J.; Du, N.; Wang, R. Increased nitrogen deposition alleviated the adverse effects of drought stress on Quercus variabilis and Quercus mongolica seedlings. Acta Physiol. Plant. 2015, 37, 107. [Google Scholar] [CrossRef]
  208. Gessler, A.; Schaub, M.; McDowell, N.G. The role of nutrients in drought-induced tree mortality and recovery. New Phytol. 2017, 214, 513–520. [Google Scholar] [CrossRef]
  209. Li, S.; Zhou, L.; Addo-Danso, S.D.; Ding, G.; Sun, M.; Wu, S.; Lin, S. Nitrogen supply enhances the physiological resistance of Chinese fir plantlets under polyethylene glycol (PEG)-induced drought stress. Sci. Rep. 2020, 10, 7509. [Google Scholar] [CrossRef]
  210. Sahay, S.; Robledo-Arratia, L.; Glowacka, K.; Gupta, M. Root NRT, NiR, AMT, GS, GOGAT and GDH expression levels reveal NO and ABA mediated drought tolerance in Brassica juncea L. Sci. Rep. 2021, 11, 7992. [Google Scholar] [CrossRef]
  211. Yang, Y.; Tang, M.; Sulpice, R.; Chen, H.; Tian, S.; Ban, Y. Arbuscular mycorrhizal fungi alter fractal dimension characteristics of Robinia pseudoacacia L. seedlings through regulating plant growth, leaf water status, photosynthesis, and nutrient concentration under drought stress. J. Plant Growth Regul. 2014, 33, 612–625. [Google Scholar] [CrossRef]
  212. Khan, A.; Tan, D.K.Y.; Afridi, M.Z.; Luo, H.; Tung, S.A.; Ajab, M.; Fahad, S. Nitrogen fertility and abiotic stresses management in cotton crop: A review. Environ. Sci. Pollut. Res. 2017, 24, 14551–14566. [Google Scholar] [CrossRef] [PubMed]
  213. Diao, M.; Ma, L.; Wang, J.; Cui, J.; Fu, A.; Liu, H.-y. Selenium Promotes the Growth and Photosynthesis of Tomato Seedlings Under Salt Stress by Enhancing Chloroplast Antioxidant Defense System. J. Plant Growth Regul. 2014, 33, 671–682. [Google Scholar] [CrossRef]
  214. Zhang, N.; Zhao, B.; Zhang, H.-J.; Weeda, S.; Yang, C.; Yang, Z.-C.; Ren, S.; Guo, Y.-D. Melatonin promotes water-stress tolerance, lateral root formation, and seed germination in cucumber (Cucumis sativus L.). J. Pineal Res. 2013, 54, 15–23. [Google Scholar] [CrossRef] [PubMed]
  215. Yousfi, S.; Márquez, A.J.; Betti, M.; Araus, J.L.; Serret, M.D. Gene expression and physiological responses to salinity and water stress of contrasting durum wheat genotypes. J. Integr. Plant Biol. 2016, 58, 48–66. [Google Scholar] [CrossRef]
  216. Teixeira, J.; Pereira, S. High salinity and drought act on an organ-dependent manner on potato glutamine synthetase expression and accumulation. Environ. Exp. Bot. 2007, 60, 121–126. [Google Scholar] [CrossRef]
  217. Cheng, L.; Wang, Y.; He, Q.; Li, H.; Zhang, X.; Zhang, F. Comparative proteomics illustrates the complexity of drought resistance mechanisms in two wheat (Triticum aestivum L.) cultivars under dehydration and rehydration. BMC Plant Biol. 2016, 16, 188. [Google Scholar] [CrossRef] [PubMed]
  218. Nagy, Z.; Németh, E.; Guóth, A.; Bona, L.; Wodala, B.; Pécsváradi, A. Metabolic indicators of drought stress tolerance in wheat: Glutamine synthetase isoenzymes and Rubisco. Plant Physiol. Biochem. 2013, 67, 48–54. [Google Scholar] [CrossRef] [PubMed]
  219. Chien, H.-F.; Lin, C.C.; Wang, J.-W.; Chen, C.T.; Kao, C.H. Changes in ammonium ion content and glutamine synthetase activity in rice leaves caused by excess cadmium are a consequence of oxidative damage. Plant Growth Regul. 2002, 36, 41–47. [Google Scholar] [CrossRef]
  220. Balestrasse, K.B.; Benavides, M.P.; Gallego, S.M.; Tomaro, M.L. Effect of cadmium stress on nitrogen metabolism in nodules and roots of soybean plants. Funct. Plant Biol. 2003, 30, 57–64. [Google Scholar] [CrossRef]
  221. Chaffei, C.; Pageau, K.; Suzuki, A.; Gouia, H.; Ghorbel, M.H.; Masclaux-Daubresse, C. Cadmium Toxicity Induced Changes in Nitrogen Management in Lycopersicon esculentum Leading to a Metabolic Safeguard Through an Amino Acid Storage Strategy. Plant Cell Physiol. 2004, 45, 1681–1693. [Google Scholar] [CrossRef]
  222. Rana, N.K.; Mohanpuria, P.; Yadav, S.K. Expression of tea cytosolic glutamine synthetase is tissue specific and induced by cadmium and salt stress. Biol. Plant. 2008, 52, 361–364. [Google Scholar] [CrossRef]
  223. Kolbert, Z.; Feigl, G.; Freschi, L.; Poór, P. Gasotransmitters in action: Nitric oxide-ethylene crosstalk during plant growth and abiotic stress responses. Antioxidants 2019, 8, 167. [Google Scholar] [CrossRef] [PubMed]
  224. Cao, Y.; Tanaka, K.; Nguyen, C.T.; Stacey, G. Extracellular ATP is a central signaling molecule in plant stress responses. Curr. Opin. Plant Biol. 2014, 20, 82–87. [Google Scholar] [CrossRef] [PubMed]
  225. Neill, S.J.; Desikan, R.; Hancock, J.T. Nitric oxide signalling in plants. New Phytol. 2003, 159, 11–35. [Google Scholar] [CrossRef] [PubMed]
  226. Chamizo-Ampudia, A.; Sanz-Luque, E.; Llamas, A.; Galvan, A.; Fernandez, E. Nitrate reductase regulates plant nitric oxide homeostasis. Trends Plant Sci. 2017, 22, 163–174. [Google Scholar] [CrossRef] [PubMed]
  227. Wilson, I.D.; Neill, S.J.; Hancock, J.T. Nitric oxide synthesis and signalling in plants. Plant Cell Environ. 2008, 31, 622–631. [Google Scholar] [CrossRef] [PubMed]
  228. Nabi, R.B.S.; Tayade, R.; Hussain, A.; Kulkarni, K.P.; Imran, Q.M.; Mun, B.-G.; Yun, B.-W. Nitric oxide regulates plant responses to drought, salinity, and heavy metal stress. Environ. Exp. Bot. 2019, 161, 120–133. [Google Scholar] [CrossRef]
  229. Pissolato, M.D.; Silveira, N.M.; Prataviera, P.J.C.; Machado, E.C.; Seabra, A.B.; Pelegrino, M.T.; Sodek, L.; Ribeiro, R.V. Enhanced Nitric Oxide Synthesis Through Nitrate Supply Improves Drought Tolerance of Sugarcane Plants. Front. Plant Sci. 2020, 11, 970. [Google Scholar] [CrossRef]
  230. Procházková, D.; Sumaira, J.; Wilhelmová, N.a.; Pavlíková, D.; Száková, J. Chapter 11—Reactive Nitrogen Species and the Role of NO in Abiotic Stress. In Emerging Technologies and Management of Crop Stress Tolerance; Ahmad, P., Rasool, S., Eds.; Academic Press: San Diego, CA, USA, 2014. [Google Scholar]
  231. Simontacchi, M.; Galatro, A.; Ramos-Artuso, F.; Santa-María, G.E. Plant Survival in a Changing Environment: The Role of Nitric Oxide in Plant Responses to Abiotic Stress. Front. Plant Sci. 2015, 6, 977. [Google Scholar] [CrossRef]
  232. Lau, S.-E.; Hamdan, M.F.; Pua, T.-L.; Saidi, N.B.; Tan, B.C. Plant Nitric Oxide Signaling under Drought Stress. Plants 2021, 10, 360. [Google Scholar] [CrossRef]
  233. Zhang, A.; Jiang, M.; Zhang, J.; Ding, H.; Xu, S.; Hu, X.; Tan, M. Nitric oxide induced by hydrogen peroxide mediates abscisic acid-induced activation of the mitogen-activated protein kinase cascade involved in antioxidant defense in maize leaves. New Phytol. 2007, 175, 36–50. [Google Scholar] [CrossRef] [PubMed]
  234. Berger, A.; Boscari, A.; Puppo, A.; Brouquisse, R. Nitrate reductases and hemoglobins control nitrogen-fixing symbiosis by regulating nitric oxide accumulation. J. Exp. Bot. 2021, 72, 873–884. [Google Scholar] [CrossRef] [PubMed]
  235. Berger, A.; Boscari, A.; Horta Araujo, N.; Maucourt, M.; Hanchi, M.; Bernillon, S.; Rolin, D.; Puppo, A.; Brouquisse, R. Plant nitrate reductases regulate nitric oxide production and nitrogen-fixing metabolism during the Medicago truncatulaSinorhizobium meliloti symbiosis. Front. Plant Sci. 2020, 11, 535004. [Google Scholar] [CrossRef] [PubMed]
  236. Wildt, J.; Kley, D.; Rockel, A.; Rockel, P.; Segschneider, H. Emission of NO from several higher plant species. J. Geophys. Res. Atmos. 1997, 102, 5919–5927. [Google Scholar] [CrossRef]
  237. Wu, Y.; Li, J.; Wang, J.; Dawuda, M.M.; Liao, W.; Meng, X.; Yuan, H.; Xie, J.; Tang, Z.; Lyu, J. Heme is involved in the exogenous ALA-promoted growth and antioxidant defense system of cucumber seedlings under salt stress. BMC Plant Biol. 2022, 22, 329. [Google Scholar] [CrossRef] [PubMed]
  238. Gan, L.; Wu, X.; Zhong, Y. Exogenously Applied Nitric Oxide Enhances the Drought Tolerance in Hulless Barley. Plant Prod. Sci. 2015, 18, 52–56. [Google Scholar] [CrossRef]
  239. Ahanger, M.A.; Qin, C.; Begum, N.; Maodong, Q.; Dong, X.X.; El-Esawi, M.; El-Sheikh, M.A.; Alatar, A.A.; Zhang, L. Nitrogen availability prevents oxidative effects of salinity on wheat growth and photosynthesis by up-regulating the antioxidants and osmolytes metabolism, and secondary metabolite accumulation. BMC Plant Biol. 2019, 19, 479. [Google Scholar] [CrossRef] [PubMed]
  240. Wang, L.; Zheng, J.; Zhou, G.; Li, J.; Qian, C.; Lin, G.; Li, Y.; Zuo, Q. Moderate nitrogen application improved salt tolerance by enhancing photosynthesis, antioxidants, and osmotic adjustment in rapeseed (Brassica napus L.). Front. Plant Sci. 2023, 14, 1196319. [Google Scholar] [CrossRef]
  241. Nazir, F.; Mahajan, M.; Khatoon, S.; Albaqami, M.; Ashfaque, F.; Chhillar, H.; Chopra, P.; Khan, M.I.R. Sustaining nitrogen dynamics: A critical aspect for improving salt tolerance in plants. Front. Plant Sci. 2023, 14, 1087946. [Google Scholar] [CrossRef]
  242. Han, M.-L.; Lv, Q.-Y.; Zhang, J.; Wang, T.; Zhang, C.-X.; Tan, R.-J.; Wang, Y.-L.; Zhong, L.-Y.; Gao, Y.-Q.; Chao, Z.-F.; et al. Decreasing nitrogen assimilation under drought stress by suppressing DST-mediated activation of Nitrate Reductase 1.2 in rice. Mol. Plant 2022, 15, 167–178. [Google Scholar] [CrossRef]
  243. Hessini, K.; Issaoui, K.; Ferchichi, S.; Saif, T.; Abdelly, C.; Siddique, K.H.M.; Cruz, C. Interactive effects of salinity and nitrogen forms on plant growth, photosynthesis and osmotic adjustment in maize. Plant Physiol. Biochem. 2019, 139, 171–178. [Google Scholar] [CrossRef] [PubMed]
  244. Liu, M.; Zhao, Y.; Liu, X.; Korpelainen, H.; Li, C. Ammonium and nitrate affect sexually different responses to salt stress in Populus cathayana. Physiol. Plant. 2022, 174, e13626. [Google Scholar] [CrossRef] [PubMed]
  245. Qu, C.; Liu, C.; Ze, Y.; Gong, X.; Hong, M.; Wang, L.; Hong, F. Inhibition of Nitrogen and Photosynthetic Carbon Assimilation of Maize Seedlings by Exposure to a Combination of Salt Stress and Potassium-Deficient Stress. Biol. Trace Elem. Res. 2011, 144, 1159–1174. [Google Scholar] [CrossRef] [PubMed]
  246. Akram, M. Effect of Nitrogen Nutrition on Solute Accumulation and Ion Contents of Maize under Sodium Chloride Stress. Commun. Soil Sci. Plant Anal. 2014, 45, 86–100. [Google Scholar] [CrossRef]
  247. Singh, M.; Singh, V.P.; Prasad, S.M. Responses of photosynthesis, nitrogen and proline metabolism to salinity stress in Solanum lycopersicum under different levels of nitrogen supplementation. Plant Physiol. Biochem. 2016, 109, 72–83. [Google Scholar] [CrossRef] [PubMed]
  248. Ji, Y.; Li, Q.; Liu, G.; Selvaraj, G.; Zheng, Z.; Zou, J.; Wei, Y. Roles of Cytosolic Glutamine Synthetases in Arabidopsis Development and Stress Responses. Plant Cell Physiol. 2019, 60, 657–671. [Google Scholar] [CrossRef] [PubMed]
  249. Yan, S.; Tang, Z.; Su, W.; Sun, W. Proteomic analysis of salt stress-responsive proteins in rice root. Proteomics 2005, 5, 235–244. [Google Scholar] [CrossRef] [PubMed]
  250. Zhao, F.G.; Qin, P. Protective effect of exogenous polyamines on root tonoplast function against salt stress in barley seedlings. Plant Growth Regul. 2004, 42, 97–103. [Google Scholar] [CrossRef]
  251. Hasanuzzaman, M.; Oku, H.; Nahar, K.; Bhuyan, M.H.M.B.; Mahmud, J.A.; Baluska, F.; Fujita, M. Nitric oxide-induced salt stress tolerance in plants: ROS metabolism, signaling, and molecular interactions. Plant Biotechnol. Rep. 2018, 12, 77–92. [Google Scholar] [CrossRef]
  252. Bhardwaj, S.; Kapoor, D.; Singh, S.; Gautam, V.; Dhanjal, D.S.; Jan, S.; Ramamurthy, P.C.; Prasad, R.; Singh, J. Nitric Oxide: A Ubiquitous Signal Molecule for Enhancing Plant Tolerance to Salinity Stress and Their Molecular Mechanisms. J. Plant Growth Regul. 2021, 40, 2329–2341. [Google Scholar] [CrossRef]
  253. Ahmad, P.; Abdel Latef, A.A.; Hashem, A.; Abd_Allah, E.F.; Gucel, S.; Tran, L.-S.P. Nitric Oxide Mitigates Salt Stress by Regulating Levels of Osmolytes and Antioxidant Enzymes in Chickpea. Front. Plant Sci. 2016, 7, 347. [Google Scholar] [CrossRef] [PubMed]
  254. Zheng, C.; Jiang, D.; Liu, F.; Dai, T.; Liu, W.; Jing, Q.; Cao, W. Exogenous nitric oxide improves seed germination in wheat against mitochondrial oxidative damage induced by high salinity. Environ. Exp. Bot. 2009, 67, 222–227. [Google Scholar] [CrossRef]
  255. Lopez, G.; Ahmadi, S.H.; Amelung, W.; Athmann, M.; Ewert, F.; Gaiser, T.; Gocke, M.I.; Kautz, T.; Postma, J.; Rachmilevitch, S.; et al. Nutrient deficiency effects on root architecture and root-to-shoot ratio in arable crops. Front. Plant Sci. 2023, 13, 1067498. [Google Scholar] [CrossRef] [PubMed]
  256. Falcioni, R.; Moriwaki, T.; Rodrigues, M.; de Oliveira, K.M.; Furlanetto, R.H.; dos Reis, A.S.; dos Santos, G.L.A.A.; Mendonça, W.A.; Crusiol, L.G.T.; Gonçalves, J.V.F.; et al. Nutrient deficiency lowers photochemical and carboxylation efficiency in tobacco. Theor. Exp. Plant Physiol. 2023, 35, 81–97. [Google Scholar] [CrossRef]
  257. Sattelmacher, B.; Addiscott, T.M. AP Whitmore and DS Powlson: Farming, Fertilizers and the Nitrate Problem; CAB International: Wallingford, UK, 1991; ISBN 0851986587. [Google Scholar]
  258. Aparicio, V.; Costa, J.L.; Zamora, M. Nitrate leaching assessment in a long-term experiment under supplementary irrigation in humid Argentina. Agric. Water Manag. 2008, 95, 1361–1372. [Google Scholar] [CrossRef]
  259. Ward, M.H. Too Much of a Good Thing? Nitrate from Nitrogen Fertilizers and Cancer. Rev. Environ. Health 2009, 24, 357–363. [Google Scholar] [CrossRef] [PubMed]
  260. Ju, X.T.; Kou, C.L.; Zhang, F.S.; Christie, P. Nitrogen balance and groundwater nitrate contamination: Comparison among three intensive cropping systems on the North China Plain. Environ. Pollut. 2006, 143, 117–125. [Google Scholar] [CrossRef]
  261. Maghanga, J.K.; Kituyi, J.L.; Kisinyo, P.O.; Ng’etich, W.K. Impact of Nitrogen Fertilizer Applications on Surface Water Nitrate Levels within a Kenyan Tea Plantation. J. Chem. 2013, 2013, 196516. [Google Scholar] [CrossRef]
  262. Bryan, N.S.; van Grinsven, H. Chapter Three—The Role of Nitrate in Human Health. In Advances in Agronomy; Sparks, D.L., Ed.; Academic Press: Cambridge, MA, USA, 2013; Volume 119, pp. 153–182. [Google Scholar]
  263. Fraser, P.; Chilvers, C.; Beral, V.; Hill, M.J. Nitrata and Human Cancer: A Review of the Evidence. Int. J. Epidemiol. 1980, 9, 3–12. [Google Scholar] [CrossRef]
  264. Knobeloch, L.; Salna, B.; Hogan, A.; Postle, J.; Anderson, H. Blue babies and nitrate-contaminated well water. Environ. Health Perspect. 2000, 108, 675–678. [Google Scholar] [CrossRef]
  265. Fischer, R.A.; Connor, D.J. Issues for cropping and agricultural science in the next 20 years. Field Crops Res. 2018, 222, 121–142. [Google Scholar] [CrossRef]
  266. Liu, Q.; Chen, X.; Wu, K.; Fu, X. Nitrogen signaling and use efficiency in plants: What’s new? Curr. Opin. Plant Biol. 2015, 27, 192–198. [Google Scholar] [CrossRef] [PubMed]
  267. Xu, G.; Fan, X.; Miller, A.J. Plant Nitrogen Assimilation and Use Efficiency. Annu. Rev. Plant Biol. 2012, 63, 153–182. [Google Scholar] [CrossRef]
  268. Motavalli, P.P.; Goyne, K.W.; Udawatta, R.P. Environmental Impacts of Enhanced-Efficiency Nitrogen Fertilizers. Crop Manag. 2008, 7, 1–15. [Google Scholar] [CrossRef]
  269. Woods, J.; Williams, A.; Hughes, J.K.; Black, M.; Murphy, R. Energy and the food system. Philos. Trans. R. Soc. B Biol. Sci. 2010, 365, 2991–3006. [Google Scholar] [CrossRef] [PubMed]
  270. Ntatsi, G.; Karkanis, A.; Yfantopoulos, D.; Pappa, V.; Konosonoka, I.H.; Travlos, I.; Bilalis, D.; Bebeli, P.; Savvas, D. Evaluation of the field performance, nitrogen fixation efficiency and competitive ability of pea landraces grown under organic and conventional farming systems. Arch. Agron. Soil Sci. 2019, 65, 294–307. [Google Scholar] [CrossRef]
  271. Van Deynze, A.; Zamora, P.; Delaux, P.-M.; Heitmann, C.; Jayaraman, D.; Rajasekar, S.; Graham, D.; Maeda, J.; Gibson, D.; Schwartz, K.D.; et al. Nitrogen fixation in a landrace of maize is supported by a mucilage-associated diazotrophic microbiota. PLoS Biol. 2018, 16, e2006352. [Google Scholar] [CrossRef] [PubMed]
  272. Barnawal, D.; Maji, D.; Bharti, N.; Chanotiya, C.S.; Kalra, A. ACC Deaminase-Containing Bacillus subtilis Reduces Stress Ethylene-Induced Damage and Improves Mycorrhizal Colonization and Rhizobial Nodulation in Trigonella foenum-graecum Under Drought Stress. J. Plant Growth Regul. 2013, 32, 809–822. [Google Scholar] [CrossRef]
  273. Kour, D.; Rana, K.L.; Yadav, A.N.; Yadav, N.; Kumar, M.; Kumar, V.; Vyas, P.; Dhaliwal, H.S.; Saxena, A.K. Microbial biofertilizers: Bioresources and eco-friendly technologies for agricultural and environmental sustainability. Biocatal. Agric. Biotechnol. 2020, 23, 101487. [Google Scholar] [CrossRef]
  274. Compant, S.; Cambon, M.C.; Vacher, C.; Mitter, B.; Samad, A.; Sessitsch, A. The plant endosphere world—Bacterial life within plants. Environ. Microbiol. 2021, 23, 1812–1829. [Google Scholar] [CrossRef]
  275. Rane, N.R.; Tapase, S.; Kanojia, A.; Watharkar, A.; Salama, E.-S.; Jang, M.; Kumar Yadav, K.; Amin, M.A.; Cabral-Pinto, M.M.S.; Jadhav, J.P.; et al. Molecular insights into plant–microbe interactions for sustainable remediation of contaminated environment. Bioresour. Technol. 2022, 344, 126246. [Google Scholar] [CrossRef] [PubMed]
  276. Dubois, O. The State of the World’s Land and Water Resources for Food and Agriculture: Managing Systems at Risk; Earthscan: Oxford, UK, 2011. [Google Scholar]
  277. Francis, B.; Aravindakumar, C.T.; Brewer, P.B.; Simon, S. Plant nutrient stress adaptation: A prospect for fertilizer limited agriculture. Environ. Exp. Bot. 2023, 213, 105431. [Google Scholar] [CrossRef]
  278. Hartemink, A.E. The definition of soil since the early 1800s. Adv. Agron. 2016, 137, 73–126. [Google Scholar]
  279. Campanhola, C.; Pandey, S. Sustainable Food and Agriculture: An Integrated Approach; Academic Press: Cambridge, MA, USA, 2018. [Google Scholar]
  280. Igiehon, N.O.; Babalola, O.O. Biofertilizers and sustainable agriculture: Exploring arbuscular mycorrhizal fungi. Appl. Microbiol. Biotechnol. 2017, 101, 4871–4881. [Google Scholar] [CrossRef] [PubMed]
  281. Bodelier, P.L. Toward understanding, managing, and protecting microbial ecosystems. Front. Microbiol. 2011, 2, 80. [Google Scholar] [CrossRef] [PubMed]
  282. Bhattarai, A.; Bhattarai, B.; Pandey, S. Variation of soil microbial population in different soil horizons. J. Microbiol. Exp. 2015, 2, 00044. [Google Scholar] [CrossRef]
  283. Xu, J.; Zhang, Y.; Zhang, P.; Trivedi, P.; Riera, N.; Wang, Y.; Liu, X.; Fan, G.; Tang, J.; Coletta-Filho, H.D. The structure and function of the global citrus rhizosphere microbiome. Nat. Commun. 2018, 9, 4894. [Google Scholar] [CrossRef]
  284. Hu, J.; Wei, Z.; Friman, V.-P.; Gu, S.-H.; Wang, X.-F.; Eisenhauer, N.; Yang, T.-J.; Ma, J.; Shen, Q.-R.; Xu, Y.-C. Probiotic diversity enhances rhizosphere microbiome function and plant disease suppression. MBio 2016, 7, e01790-16. [Google Scholar] [CrossRef]
  285. Mendes, R.; Garbeva, P.; Raaijmakers, J.M. The rhizosphere microbiome: Significance of plant beneficial, plant pathogenic, and human pathogenic microorganisms. FEMS Microbiol. Rev. 2013, 37, 634–663. [Google Scholar] [CrossRef]
  286. Mitter, E.K.; Tosi, M.; Obregón, D.; Dunfield, K.E.; Germida, J.J. Rethinking crop nutrition in times of modern microbiology: Innovative biofertilizer technologies. Front. Sustain. Food Syst. 2021, 5, 606815. [Google Scholar] [CrossRef]
  287. Nannipieri, P.; Ascher-Jenull, J.; Ceccherini, M.T.; Pietramellara, G.; Renella, G.; Schloter, M. Beyond microbial diversity for predicting soil functions: A mini review. Pedosphere 2020, 30, 5–17. [Google Scholar] [CrossRef]
  288. Johansson, J.F.; Paul, L.R.; Finlay, R.D. Microbial interactions in the mycorrhizosphere and their significance for sustainable agriculture. FEMS Microbiol. Ecol. 2004, 48, 1–13. [Google Scholar] [CrossRef] [PubMed]
  289. Saccá, M.L.; Barra Caracciolo, A.; Di Lenola, M.; Grenni, P. Ecosystem services provided by soil microorganisms. In Soil Biological Communities and Ecosystem Resilience; Springer: Cham, Switzerland, 2017; pp. 9–24. [Google Scholar]
  290. Nannipieri, P.; Ascher, J.; Ceccherini, M.; Landi, L.; Pietramellara, G.; Renella, G. Microbial diversity and soil functions. Eur. J. Soil Sci. 2003, 54, 655–670. [Google Scholar] [CrossRef]
  291. Masson-Boivin, C.; Giraud, E.; Perret, X.; Batut, J. Establishing nitrogen-fixing symbiosis with legumes: How many rhizobium recipes? Trends Microbiol. 2009, 17, 458–466. [Google Scholar] [CrossRef] [PubMed]
  292. Mendoza-Suárez, M.; Andersen, S.U.; Poole, P.S.; Sánchez-Cañizares, C. Competition, nodule occupancy, and persistence of inoculant strains: Key factors in the rhizobium-legume symbioses. Front. Plant Sci. 2021, 12, 690567. [Google Scholar] [CrossRef] [PubMed]
  293. Holland, P.L. Introduction: Reactivity of Nitrogen from the Ground to the Atmosphere. Chem. Rev. 2020, 120, 4919–4920. [Google Scholar] [CrossRef] [PubMed]
  294. Nyaga, J.W.; Njeru, E.M. Potential of native rhizobia to improve cowpea growth and production in semiarid regions of Kenya. Front. Agron. 2020, 2, 606293. [Google Scholar] [CrossRef]
  295. Yamada, Y.; Hoshino, K.-i.; Ishikawa, T. The phylogeny of acetic acid bacteria based on the partial sequences of 16S ribosomal RNA: The elevation of the subgenus Gluconoacetobacter to the generic level. Biosci. Biotechnol. Biochem. 1997, 61, 1244–1251. [Google Scholar] [CrossRef]
  296. Yan, D.; Tajima, H.; Cline, L.C.; Fong, R.Y.; Ottaviani, J.I.; Shapiro, H.Y.; Blumwald, E. Genetic modification of flavone biosynthesis in rice enhances biofilm formation of soil diazotrophic bacteria and biological nitrogen fixation. Plant Biotechnol. J. 2022, 20, 2135–2148. [Google Scholar] [CrossRef]
  297. Cesco, S.; Mimmo, T.; Tonon, G.; Tomasi, N.; Pinton, R.; Terzano, R.; Neumann, G.; Weisskopf, L.; Renella, G.; Landi, L. Plant-borne flavonoids released into the rhizosphere: Impact on soil bio-activities related to plant nutrition. A review. Biol. Fertil. Soils 2012, 48, 123–149. [Google Scholar] [CrossRef]
  298. Wang, D.; Xu, A.; Elmerich, C.; Ma, L.Z. Biofilm formation enables free-living nitrogen-fixing rhizobacteria to fix nitrogen under aerobic conditions. ISME J. 2017, 11, 1602–1613. [Google Scholar] [CrossRef] [PubMed]
  299. Flemming, H.-C.; Wingender, J.; Szewzyk, U.; Steinberg, P.; Rice, S.A.; Kjelleberg, S. Biofilms: An emergent form of bacterial life. Nat. Rev. Microbiol. 2016, 14, 563–575. [Google Scholar] [CrossRef]
  300. Young, J.M.; Leschine, S.B.; Reguera, G. Reversible control of biofilm formation by Cellulomonas spp. in response to nitrogen availability. Environ. Microbiol. 2012, 14, 594–604. [Google Scholar] [CrossRef] [PubMed]
  301. Pankievicz, V.; Irving, T.B.; Maia, L.G.; Ané, J.-M. Are we there yet? The long walk towards the development of efficient symbiotic associations between nitrogen-fixing bacteria and non-leguminous crops. BMC Biol. 2019, 17, 1–17. [Google Scholar] [CrossRef] [PubMed]
  302. Ike, A.; Sriprang, R.; Ono, H.; Murooka, Y.; Yamashita, M. Bioremediation of cadmium contaminated soil using symbiosis between leguminous plant and recombinant rhizobia with the MTL4 and the PCS genes. Chemosphere 2007, 66, 1670–1676. [Google Scholar] [CrossRef] [PubMed]
  303. Rosenblueth, M.; Ormeño-Orrillo, E.; López-López, A.; Rogel, M.A.; Reyes-Hernández, B.J.; Martínez-Romero, J.C.; Reddy, P.M.; Martínez-Romero, E. Nitrogen fixation in cereals. Front. Microbiol. 2018, 9, 1794. [Google Scholar] [CrossRef] [PubMed]
  304. Hill, S.; Kennedy, C.; Kavanagh, E.; Goldberg, R.B.; Hanau, R. Nitrogen fixation gene (nifL) involved in oxygen regulation of nitrogenase synthesis in K. peumoniae. Nature 1981, 290, 424–426. [Google Scholar] [CrossRef]
  305. Dixon, R.; Kahn, D. Genetic regulation of biological nitrogen fixation. Nat. Rev. Microbiol. 2004, 2, 621–631. [Google Scholar] [CrossRef]
  306. Burgess, B.K.; Lowe, D.J. Mechanism of molybdenum nitrogenase. Chem. Rev. 1996, 96, 2983–3012. [Google Scholar] [CrossRef]
  307. Bjornsson, R.; Delgado-Jaime, M.U.; Lima, F.A.; Sippel, D.; Schlesier, J.; Weyhermüller, T.; Einsle, O.; Neese, F.; DeBeer, S. Molybdenum L-Edge XAS Spectra of MoFe Nitrogenase. Z. Für Anorg. Und Allg. Chem. 2015, 641, 65–71. [Google Scholar] [CrossRef]
  308. Setubal, J.C.; Dos Santos, P.; Goldman, B.S.; Ertesvaåg, H.; Espin, G.; Rubio, L.M.; Valla, S.; Almeida, N.F.; Balasubramanian, D.; Cromes, L. Genome sequence of Azotobacter vinelandii, an obligate aerobe specialized to support diverse anaerobic metabolic processes. J. Bacteriol. 2009, 191, 4534–4545. [Google Scholar] [CrossRef] [PubMed]
  309. Yan, Y.; Yang, J.; Dou, Y.; Chen, M.; Ping, S.; Peng, J.; Lu, W.; Zhang, W.; Yao, Z.; Li, H. Nitrogen fixation island and rhizosphere competence traits in the genome of root-associated Pseudomonas stutzeri A1501. Proc. Natl. Acad. Sci. USA 2008, 105, 7564–7569. [Google Scholar] [CrossRef] [PubMed]
  310. Lalucat, J.; Bennasar, A.; Bosch, R.; García-Valdés, E.; Palleroni, N.J. Biology of Pseudomonas stutzeri. Microbiol. Mol. Biol. Rev. 2006, 70, 510–547. [Google Scholar] [CrossRef] [PubMed]
  311. Dingler, C.; Kuhla, J.; Wassink, H.; Oelze, J. Levels and activities of nitrogenase proteins in Azotobacter vinelandii grown at different dissolved oxygen concentrations. J. Bacteriol. 1988, 170, 2148–2152. [Google Scholar] [CrossRef] [PubMed]
  312. Desnoues, N.; Lin, M.; Guo, X.; Ma, L.; Carreño-Lopez, R.; Elmerich, C. Nitrogen fixation genetics and regulation in a Pseudomonas stutzeri strain associated with rice. Microbiology 2003, 149, 2251–2262. [Google Scholar] [CrossRef] [PubMed]
  313. Oelze, J. Respiratory protection of nitrogenase in Azotobacter species: Is a widely held hypothesis unequivocally supported by experimental evidence? FEMS Microbiol. Rev. 2000, 24, 321–333. [Google Scholar] [CrossRef] [PubMed]
  314. Curatti, L.; Brown, C.S.; Ludden, P.W.; Rubio, L.M. Genes required for rapid expression of nitrogenase activity in Azotobacter vinelandii. Proc. Natl. Acad. Sci. USA 2005, 102, 6291–6296. [Google Scholar] [CrossRef] [PubMed]
  315. Kelly, M.; Poole, R.; Yates, M.; Kennedy, C. Cloning and mutagenesis of genes encoding the cytochrome bd terminal oxidase complex in Azotobacter vinelandii: Mutants deficient in the cytochrome d complex are unable to fix nitrogen in air. J. Bacteriol. 1990, 172, 6010–6019. [Google Scholar] [CrossRef]
  316. Dedysh, S.N.; Ricke, P.; Liesack, W. NifH and NifD phylogenies: An evolutionary basis for understanding nitrogen fixation capabilities of methanotrophic bacteria. Microbiology 2004, 150, 1301–1313. [Google Scholar] [CrossRef]
  317. Raymond, J.; Siefert, J.L.; Staples, C.R.; Blankenship, R.E. The natural history of nitrogen fixation. Mol. Biol. Evol. 2004, 21, 541–554. [Google Scholar] [CrossRef]
  318. Henson, B.J.; Watson, L.E.; Barnum, S.R. The evolutionary history of nitrogen fixation, as assessed by nif D. J. Mol. Evol. 2004, 58, 390–399. [Google Scholar] [PubMed]
  319. Romanovsky, V.E.; Drozdov, D.; Oberman, N.G.; Malkova, G.; Kholodov, A.L.; Marchenko, S.; Moskalenko, N.G.; Sergeev, D.; Ukraintseva, N.; Abramov, A. Thermal state of permafrost in Russia. Permafr. Periglac. Process. 2010, 21, 136–155. [Google Scholar] [CrossRef]
  320. Christiansen, H.H.; Etzelmüller, B.; Isaksen, K.; Juliussen, H.; Farbrot, H.; Humlum, O.; Johansson, M.; Ingeman-Nielsen, T.; Kristensen, L.; Hjort, J. The thermal state of permafrost in the nordic area during the international polar year 2007–2009. Permafr. Periglac. Process. 2010, 21, 156–181. [Google Scholar] [CrossRef]
  321. Grosse, G.; Goetz, S.; McGuire, A.D.; Romanovsky, V.E.; Schuur, E.A. Changing permafrost in a warming world and feedbacks to the Earth system. Environ. Res. Lett. 2016, 11, 040201. [Google Scholar] [CrossRef]
  322. Hugelius, G.; Strauss, J.; Zubrzycki, S.; Harden, J.W.; Schuur, E.A.; Ping, C.-L.; Schirrmeister, L.; Grosse, G.; Michaelson, G.J.; Koven, C.D. Estimated stocks of circumpolar permafrost carbon with quantified uncertainty ranges and identified data gaps. Biogeosciences 2014, 11, 6573–6593. [Google Scholar] [CrossRef]
  323. Schuur, E.A.; McGuire, A.D.; Schädel, C.; Grosse, G.; Harden, J.W.; Hayes, D.J.; Hugelius, G.; Koven, C.D.; Kuhry, P.; Lawrence, D.M. Climate change and the permafrost carbon feedback. Nature 2015, 520, 171–179. [Google Scholar] [CrossRef] [PubMed]
  324. Schädel, C.; Bader, M.K.-F.; Schuur, E.A.; Biasi, C.; Bracho, R.; Čapek, P.; De Baets, S.; Diáková, K.; Ernakovich, J.; Estop-Aragones, C. Potential carbon emissions dominated by carbon dioxide from thawed permafrost soils. Nat. Clim. Chang. 2016, 6, 950–953. [Google Scholar] [CrossRef]
  325. Harden, J.W.; Koven, C.D.; Ping, C.L.; Hugelius, G.; McGuire, A.D.; Camill, P.; Jorgenson, T.; Kuhry, P.; Michaelson, G.J.; O’Donnell, J.A. Field information links permafrost carbon to physical vulnerabilities of thawing. Geophys. Res. Lett. 2012, 39, L15704. [Google Scholar] [CrossRef]
  326. Butterbach-Bahl, K.; Baggs, E.M.; Dannenmann, M.; Kiese, R.; Zechmeister-Boltenstern, S. Nitrous oxide emissions from soils: How well do we understand the processes and their controls? Philos. Trans. R. Soc. B Biol. Sci. 2013, 368, 20130122. [Google Scholar] [CrossRef]
  327. Kuypers, M.M.; Marchant, H.K.; Kartal, B. The microbial nitrogen-cycling network. Nat. Rev. Microbiol. 2018, 16, 263–276. [Google Scholar] [CrossRef]
Figure 1. Secondary structure of N transporter families in higher plants. Except for NRT1, which is represented by an experimentally solved 3D structure (PDB: 4OH3), the other four transporters are represented by their alpha fold models. Blue regions refer to the nitrogen transporter domains, and green regions refer to the rest of the protein regions.
Figure 1. Secondary structure of N transporter families in higher plants. Except for NRT1, which is represented by an experimentally solved 3D structure (PDB: 4OH3), the other four transporters are represented by their alpha fold models. Blue regions refer to the nitrogen transporter domains, and green regions refer to the rest of the protein regions.
Biomolecules 13 01443 g001
Figure 2. Overview of N metabolism pathways. (A). An illustration shows all the genes, enzymes, and substrates defined in the KEGG Pathway database regarding N metabolism. (B). The oxidation state comprises six sub-pathways: N fixation, assimilatory NO3 reduction, dissimilatory NO3 reduction, denitrification, nitrification, complete nitrification, comammox, and anammox.
Figure 2. Overview of N metabolism pathways. (A). An illustration shows all the genes, enzymes, and substrates defined in the KEGG Pathway database regarding N metabolism. (B). The oxidation state comprises six sub-pathways: N fixation, assimilatory NO3 reduction, dissimilatory NO3 reduction, denitrification, nitrification, complete nitrification, comammox, and anammox.
Biomolecules 13 01443 g002
Figure 3. The N cycle, uptake, and assimilation. The main sources of soil N are N fixation, synthetic fertilizers, and organic decomposition. N uptake in the form of NO3 or NH4+ through NRTs and AMTs, respectively. GS catalyzes glutamine synthesis from glutamate and NH4+ ions. At the same time, GOGAT catalyzes the synthesis of glutamate from glutamine and α-ketoglutarate, and thus, GS and GOGAT play central roles in the N metabolism and assimilation in plants.
Figure 3. The N cycle, uptake, and assimilation. The main sources of soil N are N fixation, synthetic fertilizers, and organic decomposition. N uptake in the form of NO3 or NH4+ through NRTs and AMTs, respectively. GS catalyzes glutamine synthesis from glutamate and NH4+ ions. At the same time, GOGAT catalyzes the synthesis of glutamate from glutamine and α-ketoglutarate, and thus, GS and GOGAT play central roles in the N metabolism and assimilation in plants.
Biomolecules 13 01443 g003
Figure 4. Adverse effects of drought stress on plant growth led to membrane damage, overproduction of ROS, and the deleterious effect on enzyme activity and other physiological processes such as photosynthesis rate. The induction of defense strategies in plants by NO through activation of antioxidant enzymes reduces the toxicity of free radicals and ROS negative impact and improves photosynthesis and membrane stability, which finally helps plants to avoid, tolerate, or escape drought conditions.
Figure 4. Adverse effects of drought stress on plant growth led to membrane damage, overproduction of ROS, and the deleterious effect on enzyme activity and other physiological processes such as photosynthesis rate. The induction of defense strategies in plants by NO through activation of antioxidant enzymes reduces the toxicity of free radicals and ROS negative impact and improves photosynthesis and membrane stability, which finally helps plants to avoid, tolerate, or escape drought conditions.
Biomolecules 13 01443 g004
Figure 5. A simplified model represents the microbial beneficial effects and the signal pathways involved in modulating the effects on N availability. This overview of plant–microbe interactions in the rhizosphere represents the bacterial colonization of the roots to enhance nutrient uptake, resulting in plant growth promotion. N source limitation also increases NH4+ and NO3 uptake activity as a promising strategy to achieve the sustainability of agriculture. In response to the nutrient deficiency, plant roots release some exudates that may recruit beneficial microbes to improve nutrient uptake. In return, they make nutrients available to plants that influence plant development. The availability of N is one of the critical elements in the process of BNF that converts atmospheric N2 to NH3. Accordingly, this model confirms that microbes might replace synthetic fertilizers in the future.
Figure 5. A simplified model represents the microbial beneficial effects and the signal pathways involved in modulating the effects on N availability. This overview of plant–microbe interactions in the rhizosphere represents the bacterial colonization of the roots to enhance nutrient uptake, resulting in plant growth promotion. N source limitation also increases NH4+ and NO3 uptake activity as a promising strategy to achieve the sustainability of agriculture. In response to the nutrient deficiency, plant roots release some exudates that may recruit beneficial microbes to improve nutrient uptake. In return, they make nutrients available to plants that influence plant development. The availability of N is one of the critical elements in the process of BNF that converts atmospheric N2 to NH3. Accordingly, this model confirms that microbes might replace synthetic fertilizers in the future.
Biomolecules 13 01443 g005
Table 1. List of genes and enzymes involved in the nitrogen cycle.
Table 1. List of genes and enzymes involved in the nitrogen cycle.
No.KEGG IDGene SymbolGene NameEC NumberReference
1K04561NORBNitric oxide reductase subunit B1.7.2.5[125,126]
2K15877CYP55Fungal nitric oxide reductase1.7.1.14[131,132]
3K00376NOSZNitrous-oxide reductase1.7.2.4[133,134,135]
4K02586NIFDNitrogenase molybdenum-iron protein alpha chain1.18.6.1[136]
5K22896VNFDVanadium-dependent nitrogenase alpha chain1.18.6.2[137]
6R05186NIFFNitrogenase1.19.6.1[138]
7K01455FORMAMIDASEFormamidase3.5.1.49[139]
8K20935HDHHydrazine dehydrogenase1.7.2.8[140]
9K20932K20932Hydrazine synthase subunit1.7.2.7[141]
10K01672CACarbonic anhydrase4.2.1.1[142,143]
11K01725CYNSCyanate lyase4.2.1.104[144]
12K00368NIRKNitrite reductase (NO-forming)1.7.2.1[145]
13K02575NRT2, NARK, NRTP, NASAMFS transporter, NNP family, nitrate/nitrite transporter------[146]
14K15576NRTA, NRTB, NRTC, NASDNitrate/nitrite transport system substrate-binding protein7.3.2.4[147]
15K00370NARG, NARZ, NXRANitrate reductase/nitrite oxidoreductase, alpha subunit1.7.5.1
1.7.99.-
[148]
16K10534NRNitrate reductase (NAD(P)H)1.7.1.1 1.7.1.2 1.7.1.3[149]
17K00367NARBFerredoxin-nitrate reductase1.7.7.2[150]
18K02567NAPANitrate reductase (cytochrome)1.9.6.1[130]
19K17877NIT-6Nitrite reductase (NAD(P)H)1.7.1.4[128]
20K00366NIRAFerredoxin-nitrite reductase1.7.7.1[151]
21K00362NIRBNitrite reductase (NADH) large subunit1.7.1.15[152]
22K03385NRFANitrite reductase (cytochrome c-552)1.7.2.2[153]
23K00459NCD2, NPDNitronate monooxygenase1.13.12.16[154]
24K19823NAONitroalkane oxidase1.7.3.1[155]
25R00143 Hydroxylamine reductase (NADH)1.7.1.10[156]
26K05601HCPHydroxylamine reductase1.7.99.1[157]
27K10535HAOHydroxylamine dehydrogenase1.7.2.6[157]
28R10230 Hydroxylamine oxidase (cytochrome)1.7.3.6[158]
29K10944PMOA-AMOAMethane/ammonia monooxygenase subunit A1.14.18.3 1.14.99.39[159]
30K00926ARCCCarbamate kinase2.7.2.2[160]
31K01948CPS1Carbamoyl-phosphate synthase (ammonia)6.3.4.16[161]
32K00260GUDB, ROCGGlutamate dehydrogenase1.4.1.2[162]
33K00261GLUD1_2, GDHAGlutamate dehydrogenase (NAD(P)+)1.4.1.3[163]
34K00262E1.4.1.4, GDHAGlutamate dehydrogenase (NADP+)1.4.1.4[164]
35K01915GLNA, GLULGlutamine synthetase6.3.1.2[165]
36K01501E3.5.5.1Nitrilase3.5.5.1[166]
37K00265GLTBGlutamate synthase (NADPH) large chain1.4.1.13[167]
38K00264GLT1Glutamate synthase (NADH)1.4.1.14[168]
39K00284GLU, GLTSGlutamate synthase (ferredoxin)1.4.7.1[169]
40K00372NASC, NASAAssimilatory nitrate reductase catalytic subunit1.7.99.-[127]
41K26139NASD, NASBNitrite reductase [NAD(P)H] large subunit1.7.1.4[127]
42K15864NIRSNitrite reductase (NO-forming)/hydroxylamine reductase1.7.2.1 1.7.99.1[170,171]
43K00531ANFGNitrogenase delta subunit1.18.6.1[120]
44K04561NORBNitric oxide reductase subunit B1.7.2.5[125,126]
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

Zayed, O.; Hewedy, O.A.; Abdelmoteleb, A.; Ali, M.; Youssef, M.S.; Roumia, A.F.; Seymour, D.; Yuan, Z.-C. Nitrogen Journey in Plants: From Uptake to Metabolism, Stress Response, and Microbe Interaction. Biomolecules 2023, 13, 1443. https://doi.org/10.3390/biom13101443

AMA Style

Zayed O, Hewedy OA, Abdelmoteleb A, Ali M, Youssef MS, Roumia AF, Seymour D, Yuan Z-C. Nitrogen Journey in Plants: From Uptake to Metabolism, Stress Response, and Microbe Interaction. Biomolecules. 2023; 13(10):1443. https://doi.org/10.3390/biom13101443

Chicago/Turabian Style

Zayed, Omar, Omar A. Hewedy, Ali Abdelmoteleb, Mohammed Ali, Mohamed S. Youssef, Ahmed F. Roumia, Danelle Seymour, and Ze-Chun Yuan. 2023. "Nitrogen Journey in Plants: From Uptake to Metabolism, Stress Response, and Microbe Interaction" Biomolecules 13, no. 10: 1443. https://doi.org/10.3390/biom13101443

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