ArticlePDF AvailableLiterature Review

Plant Growth-Promoting Soil Bacteria: Nitrogen Fixation, Phosphate Solubilization, Siderophore Production, and Other Biological Activities

Authors:

Abstract and Figures

This review covers the literature data on plant growth-promoting bacteria in soil, which can fix atmospheric nitrogen, solubilize phosphates, produce and secrete siderophores, and may exhibit several different behaviors simultaneously. We discuss perspectives for creating bacterial consortia and introducing them into the soil to increase crop productivity in agrosystems. The application of rhizosphere bacteria—which are capable of fixing nitrogen, solubilizing organic and inorganic phosphates, and secreting siderophores, as well as their consortia—has been demonstrated to meet the objectives of sustainable agriculture, such as increasing soil fertility and crop yields. The combining of plant growth-promoting bacteria with mineral fertilizers is a crucial trend that allows for a reduction in fertilizer use and is beneficial for crop production.
Content may be subject to copyright.
Citation: Timofeeva, A.M.;
Galyamova, M.R.; Sedykh, S.E. Plant
Growth-Promoting Soil Bacteria:
Nitrogen Fixation, Phosphate
Solubilization, Siderophore
Production, and Other Biological
Activities. Plants 2023,12, 4074.
https://doi.org/10.3390/
plants12244074
Academic Editors: Loretta Pace,
Christian Dimkpa and Rihab Djebaili
Received: 25 September 2023
Revised: 4 December 2023
Accepted: 4 December 2023
Published: 5 December 2023
Copyright: © 2023 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
plants
Review
Plant Growth-Promoting Soil Bacteria: Nitrogen Fixation,
Phosphate Solubilization, Siderophore Production, and Other
Biological Activities
Anna M. Timofeeva 1,2 , Maria R. Galyamova 2and Sergey E. Sedykh 1,2 ,*
1Institute of Chemical Biology and Fundamental Medicine, Siberian Branch of the Russian Academy of
Sciences, 630090 Novosibirsk, Russia; anna.m.timofeeva@gmail.com
2Faculty of Natural Sciences, Novosibirsk State University, 630090 Novosibirsk, Russia;
mgalyamova@gmail.com
*Correspondence: sedyh@niboch.nsc.ru
Abstract:
This review covers the literature data on plant growth-promoting bacteria in soil, which
can fix atmospheric nitrogen, solubilize phosphates, produce and secrete siderophores, and may
exhibit several different behaviors simultaneously. We discuss perspectives for creating bacterial
consortia and introducing them into the soil to increase crop productivity in agrosystems. The
application of rhizosphere bacteria—which are capable of fixing nitrogen, solubilizing organic and
inorganic phosphates, and secreting siderophores, as well as their consortia—has been demonstrated
to meet the objectives of sustainable agriculture, such as increasing soil fertility and crop yields. The
combining of plant growth-promoting bacteria with mineral fertilizers is a crucial trend that allows
for a reduction in fertilizer use and is beneficial for crop production.
Keywords:
plant growth-promoting bacteria; PGPB; soil bacteria; rhizosphere bacteria; phosphate
solubilization; biofertilizers; nitrogen fixation; phosphate fertilizers; siderophores; sustainable
agriculture
1. Introduction
In the second half of the twentieth century, the use of mineral fertilizers was one of the
main factors contributing to crop production development. However, unbalanced fertilizer
application has been reported to have a negative impact on crop production sustainability
and environmental safety [
1
]. Another consequence is the loss of soil microbiome diversity,
leading to reduced fertility [
2
,
3
]. A promising research avenue currently being explored
is the use of bacteria to enhance soil fertility and stimulate crop yields [
4
]. Plant growth-
promoting bacteria (PGPB) and their consortia [
5
] can naturally enhance plant growth, both
directly and indirectly, by fixing atmospheric nitrogen [
6
], synthesizing plant hormones
and siderophores [
7
], stimulating plant nutrient uptake [
8
], and suppressing pests and
pathogens [
9
,
10
]. Of particular significance is the fact that such bacteria interact with plant
roots and increase resistance to abiotic stresses [11].
This review provides a comprehensive account of the various ways in which soil
bacteria can be used to promote plant growth. We describe the nitrogen-fixing, phosphate-
solubilizing, and siderophore-producing activities of soil microorganisms. Addition-
ally, we discuss the creation of microbial consortia—including combining these with
fertilizers—that
can be introduced into soil to address the current challenges of
sustainable agriculture.
2. Nitrogen-Fixing Bacteria
Bacteria that fix atmospheric nitrogen (N
2
) live in plant tissues (e.g., tubers and roots)
and at the soil–rhizosphere interface, and can supply the significant amounts of mineral
Plants 2023,12, 4074. https://doi.org/10.3390/plants12244074 https://www.mdpi.com/journal/plants
Plants 2023,12, 4074 2 of 16
nitrogen required for plant growth [
12
,
13
]. Nitrogen fixation has been described for both
symbiotic legume bacteria and non-symbiotic soil bacteria (heterotrophic or autotrophic)
found in soil or water, or on stones or fallen leaves. Symbiotic bacteria of legume nodules
are believed to be the most critical component of the biological fixation of atmospheric ni-
trogen [
14
]. Non-symbiotic nitrogen-fixing soil bacteria are significantly underrepresented
in the literature; existing publications are mainly concerned with cereal crops such as maize,
rice, and wheat [
15
17
]. Non-symbiotic nitrogen-fixing bacteria that increase the productiv-
ity of cereal crops have been described for the following genera: Azospirillum,Azotobacter,
Beijerinckia,Burkholderia,Clostridium,Gluconacetobacter,Herbaspirillum,Methanosarcina, and
Paenibacillus [15,18,19].
Non-Symbiotic Nitrogen-Fixing Bacteria
Free-living nitrogen-fixing bacteria have been reported to use, rather than fix, nitrogen
when mineral nitrogen is available in soil [
20
,
21
]. The basic enzyme of nitrogen fixation is
nitrogenase. Its activity is sensitive to oxygen; requires metals that are part of the enzyme
subunits (such as Fe, V, and Mo); depends on ATP and reduced coenzymes; and is low in the
presence of mineral nitrogen. Free-living nitrogen-fixing bacteria can be obligate anaerobes,
facultative anaerobes, or obligate aerobes found in different environments at different
molecular oxygen concentrations. O
2
can inhibit nitrogenase and suppress N
2
fixation.
Nitrogen-fixing bacteria can avoid the potentially negative effects of O
2
by isolating the
nitrogen fixation in space, e.g., by using structures such as heterocysts, where the oxygen
concentration is kept low [22].
According to several reports, nitrogen-fixing soil bacteria supply a significant amount
of the mineral nitrogen used in agriculture [
23
], making nitrogen fixation the second most
important biogeochemical process after photosynthesis. This energy-dependent process
requires sixteen ATP molecules to fix one atmospheric nitrogen molecule [
6
]. Soil bacteria
can fix atmospheric nitrogen while exhibiting diazotrophic activity. However, they can
also convert mineral nitrogen into NO, N
2
O, and N
2
, which is considered undesirable
for agroecosystems, especially since N
2
O is a greenhouse gas. One strain of soil bacteria
tends to contain different genes for nitrogen metabolism, with their expression largely
dependent on mineral nitrogen availability in the environment. Variants of the soil bacterial
consortia have been proposed, which can perform nitrogen fixation and contribute to
nitrogen accumulation in the soils of agroecosystems [22].
All known forms of nitrogenase require Fe atoms, with most of them also containing
metals such as Mo or V [
24
]. The nifH gene, encoding nitrogenase reductase [
25
] as well
as several other genetic markers, is widely used to analyze the ability of bacteria to fix
nitrogen, as well as the distribution of nitrogen-fixing agents in communities.
The availability of cellular ATP and soluble phosphates in the environment signif-
icantly influences the process of nitrogen fixation. Additionally, this process depends
on the availability of iron ions in the environment and cellular Fe-containing cofactors
and enzymes. Nitrogen fixation, in turn, provides the nitrogen compounds necessary for
metabolic processes. Given the above, it is beyond doubt that the processes of nitrogen
fixation, phosphate solubilization, and siderophore synthesis are interrelated. The syner-
getic effect of the simultaneous introduction of a consortium of bacteria exhibiting these
activities individually, or several activities simultaneously, into the soil can stimulate plant
growth and development, meeting the goals and objectives of sustainable agriculture and
contributing to the economical and rational use of fertilizers.
3. Phosphate-Solubilizing Microorganisms
Phosphate-solubilizing bacteria [
26
] and mycorrhizal fungi [
27
] are known to in-
crease the bioavailability of phosphorus from soil to plants [
28
]. They solubilize inorganic
phosphates and mineralize insoluble organic forms of phosphorus [
29
]. Microorganisms
capable of solubilizing phosphorus have been considered in a previous review [
8
], which
describes the phosphate-solubilizing activity of bacteria belonging to the following genera:
Plants 2023,12, 4074 3 of 16
Aeromonas,Agrobacterium,Azotobacter,Bacillus,Bradyrhizobium,Burkholderia,Cyanobacteria,
Enterobacter,Erwinia,Kushneria,Micrococcus,Paenibacillus,Pseudomonas,Rhizobium,Rhodococ-
cus,Salmonella,Serratia,Serratia,Sinomonas, and Thiobacillus. Additionally, consideration is
given to fungi belonging to Achrothcium,Alternaria,Arthrobotrys,Aspergillus,Cephalosporium,
Chaetomium,Cladosporium,Cunninghamella,Curvularia,Fusarium,Glomus,Helminthosporium,
Micromonospora,Phenomiocenspora,Phenomiocenspora,Phenomycylum,Populospora,Pythium,
Rhizoctonia,Rhizopus,Saccharomyces,Schizosaccharomyces,Schwanniomyces,Sclerotium,Torula,
Trichoderma, and Yarrowia. Among all phosphate-solubilizing microorganisms, bacteria
significantly predominate, accounting for up to 50%, while fungi cover up to 0.5%. Most
phosphate-solubilizing microorganisms are found in the rhizosphere [30].
3.1. Solubilization of Inorganic Phosphorus Compounds
Phosphorus compounds in soil can be both inorganic and organic. The following
inorganic phosphate compounds have been described: apatite, strengite, and variscite; in
addition to secondary minerals, such as ferric, aluminum, and calcium phosphates [
31
].
It has been established through various studies that the primary mechanism responsible
for the mineral phosphate solubilization exhibited by bacteria is the secretion of organic
acids [
28
,
31
,
32
]. Soil bacteria have been reported to secrete organic acids, stimulating
phosphate solubilization by acidifying the environment and by chelating metal ions from
the corresponding inorganic compounds [32,33].
The solubilization efficiency also significantly depends on the strength and chemical
structure of organic acids. For example, carboxylic acids containing a single carboxyl
group are known to be less efficient than dicarboxylic or tricarboxylic acids. It has also
been demonstrated that aromatic organic acids are less active than the corresponding
aliphatic analogs. The methods of mass spectrometry analysis, gas chromatography, and
high-performance liquid chromatography [
8
] have proven that the organic acids secreted
by rhizosphere bacteria in soil are involved in phosphate solubilization. These include
acetic, adipic, butyric, citric, fumaric, glutaric, glycolic, glyconic, glyoxalic, 2-ketogluconic,
lactic, malic, malonic, oxalic, propionic, succinic, and tartaric acids [
30
], with gluconic and
2-ketogluconic acids apparently being the most important.
The secretion of organic acids by bacterial cells is associated with several metabolic
pathways, such as the direct oxidation of low molecular weight precursors in the
periplasm [34], and intracellular phosphorylation [35].
3.2. Mineralization of Organic Phosphorus Compounds
Soil’s organic phosphorus compounds can account for up to 30–50% of soil phospho-
rus. Organic phosphorus compounds are mainly found in the form of phytate (inositol
phosphate). The other organic soil phosphates described in the scientific literature include
nucleic acids; mono-, di-, and tri-esters; and phospholipids [
26
]. Xenobiotic phosphonates
are organic phosphorus compounds that can also be found in high concentrations in soil.
These include antibiotics, detergents, pesticides, flame retardants, and other compounds.
All the molecules mentioned above must be converted into soluble ionic phosphate forms
before being assimilated by plant roots [36].
Organic compounds are metabolized by enzymes secreted into the environment by
phosphate-solubilizing bacteria. Nonspecific acidic phosphatases that cleave phosphate
from the ester or phosphoanhydride bond are represented by phosphomonoesterases:
alkaline and acidic phosphatases [
37
,
38
]. The phytase enzyme has been demonstrated to
cleave phytates [
39
]. Acid phosphatase activity has been observed in Pseudomonas fluo-
rescens [
40
,
41
], Burkholderia cepacia [
42
], Enterobacter aerogenes,Enterobacter cloacae,Citrobacter
freundi,Proteus mirabalis and Serratia marcenscens [
43
], and Klebsiella aerogenes [
44
]. Addition-
ally, phytase activity has been demonstrated for Bacillus subtilis,Pseudomonas putida, and
Pseudomonas mendocina [
45
]. Interestingly, the secretion of phosphatases by soil bacteria
significantly depends on both the free phosphate already available in the soil and the
availability of inorganic nitrogen [
46
,
47
], indicating a close relationship between nitro-
Plants 2023,12, 4074 4 of 16
gen fixation and phosphate solubilization processes, and possible synergism of bacterial
consortia combining these activities.
4. Siderophore-Secreting Microorganisms
Fe
2+
ions are involved in various biochemical intracellular processes in plants, such
as atmospheric nitrogen fixation (see Section 1) and photosynthesis [
48
]. Although iron is
one of the most common elements in the Earth’s crust, its availability for plant root uptake
is extremely low due to the rapid oxidation of Fe
2+
ions to Fe
3+
ions in the environment,
making them insoluble [49].
Bacteria and other microorganisms are known to secrete siderophores—organic com-
pounds with a molecular mass of up to 1.5 kDa and a high affinity for Fe
3+
[
50
52
].
Siderophores secreted by rhizosphere microorganisms promote the conversion of Fe
2+
ions into a form available to plant roots [53,54], favoring plant growth and development.
Siderophore-producing bacteria have been described in twenty genera [
7
], includ-
ing Azospirillum,Azotobacter,Bacillus,Dickeya,Klebsiella,Nocardia,Paenibacillus,Pantoea,
Pseudomonas,Serratia,Streptomyces, and others. A number of genera from this list are
also known to contain atmospheric nitrogen-fixing and phosphate-solubilizing bacteria.
Some bacteria have been reported to exhibit multiple activities that promote plant growth
and development.
It should be noted that siderophores are not secreted only by soil bacteria. They can
also be produced by human pathogenic bacteria [55], fungi [56,57], and oomycetes [58].
4.1. Siderophores Produced by Soil Bacteria
The siderophore structure comprises an iron atom surrounded and coordinated by
oxygen atoms. The octahedral configuration that is most common in bacterial siderophores
contributes to the stabilization of Fe
3+
ions [
59
]. Fe
3+
ions can be coordinated by various
functional structural groups, with the most common being catecholates, hydroxamates,
α
-hydroxycarboxylates, and combinations of nitrogen-containing heterocycles, pheno-
lates, and carboxylates [
60
]. Siderophores are classified by structure and chemical na-
ture as catecholate, hydroxamate, carboxylate, and phenolate siderophores [
61
,
62
]. Some
siderophores are classified as mixed, due to their structure corresponding to two or three
classes simultaneously.
The structure of hydroxamate siderophores includes a functional group C(=O)N-
(OH)R, with R being an amino acid or its derivative. This group contains two oxygen
atoms forming a ligand with Fe
2+
ions [
61
]. Hydroxamate schizokinen siderophores have
been described in Bacillus megaterium [
63
]. The Rhizobium leguminosarum strain has also
been found to produce a schizokinen siderophore [
64
]. Other hydroxamate siderophores
have been described in Pantoea vagans [
65
], Rhizobium meliloti [
66
], R. leguminosarum, and
R. phaseoli [67,68].
Catecholate siderophores contain Fe3+ ions bound to catecholate or hydroxyl groups.
The structure of the resulting complex is also octahedral [
69
]. Catecholate siderophores
are most commonly derived from either 2,3-dihydroxybenzoic acid or salicylic acid [
70
].
Catecholate siderophores have been described in bacteria of the genus Azospirillum, in-
cluding A. brasilense [
71
], A. lipoferum [
72
], and A. vinelandii, which are known to produce
four different catecholate siderophores [
61
,
73
,
74
]. Other siderophores have been described
in Rhizobium leguminosarum [
64
], R. radiobacter (agrobactin) [
75
,
76
], Bacillus subtilis (itoic
acid) [77], and B. thuringiensis (bacillibactin) [78].
Siderophores containing several Fe
2+
chelating groups are categorized as the mixed
type. Pseudomonas strains secreting pyoverdines have been described [
79
,
80
], with their
structure comprising a quinoline chromophore, a peptide, and a dicarboxylic acid or its
amide attached to the chromophore [81].
For various strains of Pseudomonas fluorescens, several pyoverdins [
82
], enantio-
pyochelin [
83
], quinolobactin [
84
], ornichorrugatin, and prepseudomonin have been de-
scribed [
85
]. Pantoea eucalypti has been shown to secrete pyoverdine- and pyochelin-like
Plants 2023,12, 4074 5 of 16
siderophores in the alkaline medium [
86
]. The pyoverdine siderophore azotobactin has
been described in the A. vinelandii strain [87].
It is noteworthy that Fe ions possess a higher affinity for bacterial siderophores com-
pared to fungal siderophores, providing the former with antifungal properties. To date, the
antifungal properties of the pyoverdine of Pseudomonas sp. have been investigated most
thoroughly [
88
]. In summary, not only can pyoverdines favorably affect plant growth, but
they can also exhibit antifungal properties and control relevant phytopathogens [89].
4.2. Life Cycle of Siderophores
Siderophores are synthesized in the cytoplasm and peroxisomes [
49
,
90
]. Next,
siderophore molecules are secreted into the extracellular space. Following the binding of
siderophores to Fe
3+
ions, the complexes of iron ions and siderophores are transported
into the periplasm by specific receptors. This transport mechanism—as opposed to a
concentration gradient—is energy dependent, further confirming that the processes of ATP
synthesis, atmospheric nitrogen fixation, and phosphate solubilization in the bacterial cell
are interrelated.
The availability of iron ions for intracellular processes is due to two mechanisms: the
reduction of Fe
3+
ions to Fe
2+
ions and the dissociation of the iron–siderophore complex,
or the binding of ions to a higher-affinity acceptor [
59
]. Siderophore-interacting proteins,
siderophore reductase, and enzymes that hydrolyze iron–siderophore complexes have
been characterized [
49
]. It is apparent that the reduction of an iron ion in the siderophore
complex is much more advantageous for the cell, allowing the siderophore molecule to be
reused, which is unfeasible when siderophores are hydrolyzed by hydrolases.
Siderophores secreted by rhizosphere bacteria provide iron to plant roots and stimulate
plant growth whenever the bioavailability of iron ions is low. Two mechanisms by which
plants obtain iron ions from siderophores secreted by microorganisms have been described.
The first mechanism involves the transport of siderophore complexes with iron ions to the
apoplast, leading to siderophore reduction and Fe
2+
ion delivery to the root. The second
mechanism allows iron ions to be exchanged between bacterial siderophores and plant
siderophores [91].
5. Isolation and Biochemical Analysis of Soil PGPB
Since it is impractical to conduct
in vivo
experiments with numerous bacterial isolates,
the procedure of soil PGPB analysis is first directed toward the analysis of useful traits.
Thus, PGPB are inoculated into plants, and the strains showing the best results are chosen,
as described in [
92
]. Almost all methods for testing bacterial activity involve the primary
isolation of pure cultures of microorganisms followed by inoculation into different media,
with the corresponding activity assessed by growth or changes in color/transparency
around the colony. Figure 1illustrates the beneficial effects of soil PGPB on the soil’s
biological activities towards promoting plant growth and development.
5.1. Isolation of Nitrogen-Fixing Bacteria
A 10 g soil sample is usually used to isolate soil bacteria. The procedure involves
sample homogenization with 90 mL of sterile water, and using serial dilutions (1:10) up to
10
6
. One milliliter of the resulting solution is applied to a Petri dish containing nutrient
agar and incubated for 2–5 days at 28
C [
93
]. After incubation, colonies of soil bacteria are
dissolved in sterile saline.
To isolate diazotrophic (nitrogen-fixing) bacteria, diluted soil bacteria samples are
inoculated into nitrogen-free semisolid media (Ashby’s Mannitol Agar or various other
compositions) and incubated for up to seven days at 28
C. After the incubation, samples
in Petri dishes—with the media covered with colorless or colored colonies, plaque, or
film—are considered positive for the presence of nitrogen-fixing bacteria. A single colony
or bacterial biomass is taken from the surface of each culture medium using a sterile
loop and used to inoculate fresh N-free semisolid media of the same type. This step
Plants 2023,12, 4074 6 of 16
is usually repeated two or three times to ensure that the isolates grow in a nitrogen-
free culture medium. At each stage, a microscopic or molecular biological (PCR, 16S
sequencing) control is used to confirm that a pure strain of bacteria is transferred to the next
Petri dish [94].
Plants 2023, 12, x FOR PEER REVIEW 6 of 17
Figure 1. Different soil PGPB influence plant growth and development in various beneficial ways.
These include fixation of atmospheric nitrogen, solubilization of phosphates and potassium, secre-
tion of siderophores and phytohormones, and activity against plant pathogens. Moreover, bacterial
consortia provide cumulative effects, favoring various aspects of plant viability.
5.1. Isolation of Nitrogen-Fixing Bacteria
A 10 g soil sample is usually used to isolate soil bacteria. The procedure involves
sample homogenization with 90 mL of sterile water, and using serial dilutions (1:10) up
to 10−6. One milliliter of the resulting solution is applied to a Petri dish containing nutrient
agar and incubated for 25 days at 28 °C [93]. After incubation, colonies of soil bacteria
are dissolved in sterile saline.
To isolate diazotrophic (nitrogen-fixing) bacteria, diluted soil bacteria samples are
inoculated into nitrogen-free semisolid media (Ashbys Mannitol Agar or various other
compositions) and incubated for up to seven days at 28 °C. After the incubation, samples
in Petri disheswith the media covered with colorless or colored colonies, plaque, or
filmare considered positive for the presence of nitrogen-fixing bacteria. A single colony
or bacterial biomass is taken from the surface of each culture medium using a sterile loop
and used to inoculate fresh N-free semisolid media of the same type. This step is usually
repeated two or three times to ensure that the isolates grow in a nitrogen-free culture me-
dium. At each stage, a microscopic or molecular biological (PCR, 16S sequencing) control
is used to confirm that a pure strain of bacteria is transferred to the next Petri dish [94].
5.2. Phenotypic Characteristics and Genotyping of Bacteria
Bacterial colonies are characterized by their morphological characteristics, such as
cell and colony shape, color, elevation and opacity, and Gram staining [95]. To identify
the genus (and species) of the strains, sequencing of the 16S rRNA gene is commonly used.
This gene is used due to specific characteristics such as size, ubiquity among bacteria, and
low rate of evolution. Bacterial genomic DNA is isolated [96]. Partial amplification of the
16S rRNA gene for Sanger sequencing is carried out using 27F (5-AGAGTTT-
GATCCTGGCTCAG-3) and 1492R (5-GGTTACCTTGTTACGACTT-3) primers. Maxi-
mum likelihood phylogenetic reconstruction of 16S rRNA genes allows strain classifica-
tion to the genus or species level.
5.3. Metabolic Characterization of Plant Growth Promotion
Metabolic characterization of a strain, in the sense of biochemical activities associated
with plant growth promotion, involves growing bacteria on Petri dishes or 96-well plates
Figure 1.
Different soil PGPB influence plant growth and development in various beneficial ways.
These include fixation of atmospheric nitrogen, solubilization of phosphates and potassium, secretion
of siderophores and phytohormones, and activity against plant pathogens. Moreover, bacterial
consortia provide cumulative effects, favoring various aspects of plant viability.
5.2. Phenotypic Characteristics and Genotyping of Bacteria
Bacterial colonies are characterized by their morphological characteristics, such as cell
and colony shape, color, elevation and opacity, and Gram staining [
95
]. To identify the genus
(and species) of the strains, sequencing of the 16S rRNA gene is commonly used. This gene
is used due to specific characteristics such as size, ubiquity among bacteria, and low rate of
evolution. Bacterial genomic DNA is isolated [
96
]. Partial amplification of the 16S rRNA
gene for Sanger sequencing is carried out using 27F (5
0
-AGAGTTTGATCCTGGCTCAG-3
0
)
and 1492R (5
0
-GGTTACCTTGTTACGACTT-3
0
) primers. Maximum likelihood phylogenetic
reconstruction of 16S rRNA genes allows strain classification to the genus or species level.
5.3. Metabolic Characterization of Plant Growth Promotion
Metabolic characterization of a strain, in the sense of biochemical activities associated
with plant growth promotion, involves growing bacteria on Petri dishes or 96-well plates
with different media, or assessing the production of bioactive compounds secreted by PGPB
through analytical methods.
5.3.1. Solubilization of Phosphates and Potassium
Phosphate solubilizing activity can be determined by growing bacterial isolates in a
modified agar medium. This medium contains a 2% (w/v) inorganic source of phosphate,
i.e., tricalcium phosphate (Ca
3
(PO
4
)
2
), referred to as Pi. Alternatively, lecithin is used as
the organic phosphate source, and this medium is called Po. The cultivation is conducted
at a temperature of 30
C for seven days. Clear zones forming around bacterial colonies
indicate that specific bacterial isolates are solubilizing phosphate [93,94].
The ability of bacteria to solubilize potassium is assessed with agar containing mica
powder (0.2%) as the single source of potassium. Bacterial cultures that form a clear zone
are considered K solubilizers [97,98].
These methods make it possible to quantify the biochemical characteristics of mi-
croorganisms. The ability of isolates to dissolve phosphorus is assessed by measuring
Plants 2023,12, 4074 7 of 16
the translucent halo around the colony. The phosphate solubilization index (PSI) and
potassium solubilization index (KSI) are determined by calculation, using the equation
in which the diameter of the clear zone is added to the colony diameter and the sum is
divided by the colony diameter [98,99].
5.3.2. Siderophore Production
PGPB isolates are inoculated onto chromazurol S (CAS) agar plates to screen for
siderophore-producing activity. The formation of an orange halo around the bacterial
colony indicates the production of siderophores [
100
]. Solid agar medium CAS can also
be used to qualitatively determine siderophore production. Chromazurol S is a complex
compound containing iron ions. The siderophores are strong Fe-chelators, and the removal
of iron from the complex changes its color from blue to orange.
Based on this method, a quantitative assessment of the concentration of siderophores
in the supernatant can be performed. This can be done by mixing the supernatant with the
CAS reagent (1:1), and, after 20 min, by measuring the optical density at 630 nm [
101
]. The
chromazurol assay does not reveal the type of bacterial siderophores. However, there are
at least three other identification methods. The Arnow test [
102
] allows the presence of
catechol groups in siderophores to be identified. The Csáky assay [103] is used to identify
the presence of hydroxamate-type siderophores. The Shenker test can detect carboxylate
siderophores [104].
5.3.3. Analysis of Phytohormone Production
Auxin, ethylene, abscisic acid, cytokinin, and gibberellin are prominent classes of
phytohormones, with each playing a distinct role in plant growth and development. PGPB
can supply plants with significant amounts of auxin, resulting in modifications to the
development of root systems [
103
,
104
]. These changes occur through the synergistic action
of exogenous and endogenous auxin [
105
]. In addition, enzymes produced by plant
growth-promoting bacteria, such as 1-aminocyclopropane-1-carboxylate (ACC) deaminase,
indirectly contribute to architectural and functional modifications of the root [106].
Among the phytohormones produced by PGPB, auxins and indolyl-3-acetic acid (IAA)
are the most studied [
107
,
108
]. The main pathway of their biosynthesis is carried out
through tryptophan as a precursor [
109
], and a number of papers show the importance
of bacterial IAA in promoting plant growth [
110
]. Several methods are used to determine
IAA. One method is the determination of IAA using gas–liquid chromatography with a
mass-selective detector (emission 360 nm, excitation 282 nm). This method requires the use
of expensive equipment and complex sample preparation. The second method for deter-
mining IAA is the interaction of IAA with Salkovsky’s reagent (2 mL of 0.5 M solution of
FeCl
3
in 100 mL of 37% HClO
4
), which leads to a colored compound. Inoculants are grown
in R2A broth containing 2 mM L-tryptophan, an IAA precursor. The supernatant is mixed
with Salkovsky’s reagent, and the concentration of IAA is assessed spectrophotometrically
via the optical density of the colored product at 530 nm [111113].
The enzyme 1-aminocyclopropane-1-carboxylate deaminase is responsible for ethylene
synthesis and indirectly contributes to architectural and functional modifications of the
root [
114
]. Quantitative analysis of deaminase activity is based on the ability of microor-
ganisms to assimilate 1-aminocyclopropane-1-carboxylate as the sole nitrogen source [
115
].
Bacteria are spot-inoculated into nitrogen-free saline DF (Dworkin and Foster) media on
plates supplemented with ACC, and incubated at 30
C for 5 days. Growth of isolates on
ACC-supplemented plates demonstrates deaminase activity.
5.4. Analysis of Antagonistic Activity against Phytopathogens and Competition between Strains
A test for antifungal activity is used to analyze the antagonistic activity of bacterial
strains [
116
]. Bacterial strains are evenly distributed over a Petri dish containing a potato
dextrose agar medium. The plates are incubated at 4
C for 2 h to promote the diffusion of
bacteria into the medium, and then a 6 mm diameter fungal disk sample is placed in the
Plants 2023,12, 4074 8 of 16
center of each Petri dish. The plates are incubated under dark conditions at a temperature
of 25
C for a period of seven days. Calculation of the percentage inhibition of radial
growth involves subtracting the Rs value (radius of fungus in the presence of bacteria) from
the Rc value (radius of fungus control) and dividing the remainder by Rc, according to the
formula [
116
]. Inhibition of the growth of the phytopathogen indicates the biocontrol of
the bacterial strain studied.
A cross-test is used to assess antagonistic activity between bacterial strains [
113
]. The
first strain is applied to the agar medium in perpendicular streaks and grown at 30
C for
3 days. The second strain is then applied with strokes at an angle of approximately 90
,
extending outward from the emerging colonies of the first strain. Both strains of bacteria
are incubated at 30
C for another 3 days. Colony lines and zones of inhibition that arise
at the intersection of strains are visualized. Strain compatibility is characterized by the
absence of inhibition zones at the intersection of two colonies. The incompatibility is caused
by competition between bacteria.
5.5. Analysis of Plant Growth Promotion
Sterile plant seeds are germinated and grown under sterile conditions at 22
C with
a photoperiod of 16 h of light and 8 h of darkness [
117
]. Five- to ten-day-old seedlings
are inoculated with a bacterial culture, with uninoculated plants used as controls. After
inoculation, photographs of the plants are taken every two days to determine the length
of the primary root and the number and density of lateral roots (density is the number of
lateral roots or the length of the primary root). The seedlings are weighed and dried at 80
C
in paper bags until a constant weight is reached. Wet and dry biomass are determined.
5.6. Bacterial Genes Promoting Plant Growth
Several bacterial genes are already known to be involved in the beneficial effects
observed in plants, such as (i) genes involved in atmospheric nitrogen fixation (Nif genes,
which encode the nitrogenase complex and other regulatory proteins); (ii) genes responsible
for the formation of legume nodules (nod); (iii) genes controlling pathogens (chi genes,
which produce chitinases, and sfp genes, which produce surfactins); (iv) genes involved in
the production of phytohormones (acdS, which encodes the production of ACC-deaminase,
improving stress resistance by reducing ethylene levels in the plant, and ipdC/ppdC, which
are involved in the production of indoleacetic acid); (v) genes involved in vitamin produc-
tion (pqq, which encodes pyrroloquinoline quinone); and (vi) genes involved in nutrient
mobilization (bgl/ybg genes, which are involved in phosphate solubilization, and rhb genes,
which encode siderophore production).
The nifHDK,nifDK,nifK, and nifD genes predict nitrogen fixation ability in bacterial
genomes [
118
]. Phytase is one of the main enzymes produced by phosphate-solubilizing
bacteria during the mineralization of organic phosphorus. This enzyme releases phos-
phorus from organic materials in the soil, which is stored as phytate [
39
]. Gluconic acid
secretion is the best-characterized phosphate solubilization mechanism, which is medi-
ated by pyrroloquinoline quinone (PQQ)-dependent glucose dehydrogenase (GDH) [
119
].
Analysis of the genes (pqqA,pqqE,pqqAD, and pqqD) encoding the cofactor PQQ is used
as a proxy for phosphate solubilization. Auxin, ethylene, abscisic acid, cytokinin, and gib-
berellin are well-known classes of phytohormones. PGPB can increase plant productivity
by causing changes in plant roots via auxin production, which acts synergistically with
endogenous auxin and alters the development of the plant root system [
120
]. Indole-3-
pyruvate decarboxylase (IPDC) is a key enzyme in synthesizing heteroauxin, the most
important auxin encoded by the ipdC gene.
Specific or universal primer pairs and Sanger sequencing, or NGS in the case of soil
consortia, are used to analyze the genes mentioned above.
Plants 2023,12, 4074 9 of 16
5.7. Environmental Factors That Can Affect PGPB and Increase or Decrease the Synthesis of Plant
Growth-Promoting Substances
Various environmental factors are known to affect the production of plant growth-
promoting substances by PGPB. These factors include interaction with other microorgan-
isms in the soil, climate zone soil types, soil physicochemical properties, and environmental
conditions [
32
]. For example, PGPB in soils exposed to extreme environmental conditions,
such as saline–alkaline soils, soils with high nutrient deficiencies, or soils from extreme
temperature environments, tend to solubilize more phosphate and potassium than bacteria
in soils from more moderate conditions [28].
The PGPB activity is also affected by soil composition. The growth rate of microorgan-
isms was found to be related to the concentration of soluble phosphate [
121
]. Additionally,
the growth rate of the Pseudomonas aeruginosa strain was estimated to be 25 times greater
under phosphate excess than under phosphate deficiency [
122
]. The PGPB-mediated min-
eral phosphate solubilization activity was found to be inhibited by soluble phosphate via
a negative feedback mechanism [
123
]. In this way, the phosphate-solubilizing activity of
bacteria is induced by low levels of exogenous soluble phosphate and inhibited by high
levels of exogenous soluble phosphate [33].
Various abiotic stresses can result in higher production of various growth-promoting
factors. For example, drought and/or salt stress can increase the sensitivity of many plants
to various phytopathogens, frequently by reducing the ability of the plant to effectively
attack the pathogen. Under unfavorable conditions, PGPB were found to increase the secre-
tion of the phytohormones abscisic acid, salicylic acid, and ethylene, which are responsible
for activating the signaling cascade of various genes involved in salt tolerance [124].
Figure 2provides a schematic representation of the environmental factors that affect
PGPB and the secretion of substances that enhance plant growth.
Plants 2023, 12, x FOR PEER REVIEW 9 of 17
The nifHDK, nifDK, nifK, and nifD genes predict nitrogen fixation ability in bacterial
genomes [118]. Phytase is one of the main enzymes produced by phosphate-solubilizing
bacteria during the mineralization of organic phosphorus. This enzyme releases phospho-
rus from organic materials in the soil, which is stored as phytate [39]. Gluconic acid secre-
tion is the best-characterized phosphate solubilization mechanism, which is mediated by
pyrroloquinoline quinone (PQQ)-dependent glucose dehydrogenase (GDH) [119]. Anal-
ysis of the genes (pqqA, pqqE, pqqAD, and pqqD) encoding the cofactor PQQ is used as a
proxy for phosphate solubilization. Auxin, ethylene, abscisic acid, cytokinin, and gibber-
ellin are well-known classes of phytohormones. PGPB can increase plant productivity by
causing changes in plant roots via auxin production, which acts synergistically with en-
dogenous auxin and alters the development of the plant root system [120]. Indole-3-py-
ruvate decarboxylase (IPDC) is a key enzyme in synthesizing heteroauxin, the most im-
portant auxin encoded by the ipdC gene.
Specific or universal primer pairs and Sanger sequencing, or NGS in the case of soil
consortia, are used to analyze the genes mentioned above.
5.7. Environmental Factors That Can Affect PGPB and Increase or Decrease the Synthesis of
Plant Growth-Promoting Substances
Various environmental factors are known to affect the production of plant growth-
promoting substances by PGPB. These factors include interaction with other microorgan-
isms in the soil, climate zone soil types, soil physicochemical properties, and environmen-
tal conditions [32]. For example, PGPB in soils exposed to extreme environmental condi-
tions, such as salinealkaline soils, soils with high nutrient deficiencies, or soils from ex-
treme temperature environments, tend to solubilize more phosphate and potassium than
bacteria in soils from more moderate conditions [28].
The PGPB activity is also affected by soil composition. The growth rate of microor-
ganisms was found to be related to the concentration of soluble phosphate [121]. Addi-
tionally, the growth rate of the Pseudomonas aeruginosa strain was estimated to be 25 times
greater under phosphate excess than under phosphate deficiency [122]. The PGPB-medi-
ated mineral phosphate solubilization activity was found to be inhibited by soluble phos-
phate via a negative feedback mechanism [123]. In this way, the phosphate-solubilizing
activity of bacteria is induced by low levels of exogenous soluble phosphate and inhibited
by high levels of exogenous soluble phosphate [33].
Various abiotic stresses can result in higher production of various growth-promoting
factors. For example, drought and/or salt stress can increase the sensitivity of many plants
to various phytopathogens, frequently by reducing the ability of the plant to effectively
attack the pathogen. Under unfavorable conditions, PGPB were found to increase the se-
cretion of the phytohormones abscisic acid, salicylic acid, and ethylene, which are respon-
sible for activating the signaling cascade of various genes involved in salt tolerance [124].
Figure 2 provides a schematic representation of the environmental factors that affect
PGPB and the secretion of substances that enhance plant growth.
Figure 2.
Environmental factors that affect the secretion of plant growth-favoring substances by
soil bacteria.
6. Consortia of Soil PGPB
The soil microbiome is a complex community of millions of species with billions of
possible interactions between them. Given that no soil bacterial species exist in isolation,
these interactions result in new properties that are not characteristic of individual genera,
species, and/or strains of bacteria [
5
]. Consortia formed through the interaction of several
bacteria are promising for agricultural applications [
125
,
126
] compared to single-species
inoculants [
127
]. However, it is essential for the species within a consortium not to be
antagonistic to each other, and to be able to occupy a wider range of ecological niches
than they would if they were isolated [
128
]. On the one hand, a more diverse microbial
consortium may provide a greater number of functions beneficial to plants [
129
]. On the
other hand, some strains of PGPB may simultaneously exhibit several activities favorable to
plant growth. For example, several publications report that nitrogen-fixing bacteria exhibit
phosphate-solubilizing and other activities. Alternatively, it is not known how actively a
single bacterial strain can simultaneously exhibit several traits favorable for plant growth
Plants 2023,12, 4074 10 of 16
and development. Whatever the case, the application of microbial consortia can provide
several plant-beneficial functions to the soil, to be performed simultaneously by different
consortium members due to their mutual complementarity.
The application of a mixture of Bacillus,Pseudomonas, and Streptomyces isolated from
soil and exhibiting phosphate-solubilizing activity has been shown to provide better re-
sults in accelerating the linear growth and root growth of plants compared to the use of
individual strains. Another example is the inoculation of wheat seeds with Pseudomonas
fluorescens Ms-01 and Azosprillum brasilense DSM1690 bacteria, which solubilize phosphates
and secrete auxins that strongly increase root and shoot biomass compared to the appli-
cation of the strains separately [
130
]. The consortium of Bacillus firmus,Cellulosimicrobium
cellulans, and Pseudomonas aeruginosa has been demonstrated to increase the root length
and linear growth of amaranth plants more effectively than when applying the strains one
by one or in pairs [
131
]. In recent years, an increasing number of scientific papers have
focused on the application of bacterial consortia with a positive effect on plant growth and
development, corroborating their unequivocal relevance and significance.
Combined Application of Bacteria and Phosphate Fertilizers
The coapplication of phosphate-solubilizing bacteria and some phosphate fertilizers
has been demonstrated to produce a synergistic effect and increase agronomic efficiency,
such as when using phosphorites (rock phosphates) [
132
134
]. Inoculation with phosphate-
solubilizing bacteria has been found to significantly improve corn growth when applied
with various inorganic and organic fertilizers: manure, poultry manure, superphosphate,
and phosphate [
135
]. Inoculating wheat plants with Pseudomonas sp. or Enterobacter
sp. and applying diammonium phosphate has been proven effective [
131
]. Enterobacter
sp. has been reported to increase phosphorite solubilization from 17.5% [
136
] to 27%
compared to the release from phosphorites (4%) alone [
137
]. A combination of a one-time
inoculation of wheat plants with five different strains of Pseudomonas (P. plecoglossicida,
P. reinekei,P. koreensis,P. japonica, and P. frederiksbergensis) and phosphate application was
found to increase phosphate availability, nutrient availability, and chlorophyll content,
and to improve the morphological characteristics associated with a higher phosphorus
uptake [
138
]. A consortium of three strains, P. corrugata,P. koreensis, and P. frederiksbergensis,
was found to enhance the plant growth of alfalfa (M. truncatula) when coapplied with
phosphate fertilizers [139].
Phosphate-solubilizing soil bacteria are known not only to increase the efficiency
of mineral phosphate fertilizers, but also to have beneficial effects when applied with
organic fertilizers. For example, the application of a mixture of compost (biogas production
residues) coupled with fossil phosphate-solubilizing bacteria Bacillus sp. was reported to
enhance the availability of free phosphate in soil, increase the soil’s organic matter content,
and boost the population of phosphate-solubilizing bacteria [140].
To summarize, the inoculation of crops with phosphate-solubilizing bacteria combined
with mineral or organic fertilizers is a promising integrated strategy for increasing soluble
phosphate availability, improving the agronomic efficiency of phosphate fertilizer usage,
increasing soil fertility, and achieving sustainable agriculture goals.
7. Conclusions
Recent studies have shown that the application of PGPB as a biofertilizer is of great
importance for the sustainable management of soil resources. The quality and nutritional
value of agricultural products are of great significance to consumers. Further research in
this area is necessary to provide farmers with timely information on the biological activ-
ity of PGPB. Atmospheric nitrogen fixation, phosphate solubilization, and increased iron
bioavailability in soil due to soil bacteria all contribute to sustainable agricultural develop-
ment, particularly by increasing soil fertility and crop yields. The usage of soil bacteria and
their consortia for inoculation into plant seeds, as well as for coapplication with mineral fer-
tilizers, may eventually contribute to developing a new generation of farming technologies
Plants 2023,12, 4074 11 of 16
with lower costs and reduced negative effects on the environment. Analyzing the molecular
mechanisms of nitrogen fixation, phosphate solubilization, and siderophore synthesis,
as well as identifying bacterial strains that can simultaneously exhibit several different
activities, are necessary steps to create commercial bacterial preparations for enhancing
plant growth in crops. One area that demands special consideration is the experimental
testing of both individual strains and their consortia across various climatic zones, diverse
soil types, and different crops. The authors of this review find such experimental research
highly relevant due to the potential ineffectiveness of commercial preparations.
Author Contributions:
Conceptualization, A.M.T., M.R.G. and S.E.S.; writing—original draft prepa-
ration, A.M.T.; writing—review and editing, S.E.S.; visualization, A.M.T.; funding acquisition, S.E.S.
All authors have read and agreed to the published version of the manuscript.
Funding:
This work was supported by the Ministry of Science and Higher Education of the Russian
Federation, agreement No. 075-15-2021-1085.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
References
1.
Nelson, A.R.L.E.; Ravichandran, K.; Antony, U. The impact of the Green Revolution on indigenous crops of India. J. Ethn. Foods
2019,6, 8. [CrossRef]
2.
Fernández-Romero, M.; Parras-Alcántara, L.; Lozano-García, B.; Clark, J.; Collins, C. Soil Quality Assessment Based on Carbon
Stratification Index in Different Olive Grove Management Practices in Mediterranean Areas. Catena
2016
,137, 449–458. [CrossRef]
3.
Patra, S.; Mishra, P.; Mahapatra, S.C.; Mithun, S.K. Modelling Impacts of Chemical Fertilizer on Agricultural Production: A Case
Study on Hooghly District, West Bengal, India. Model. Earth Syst. Environ. 2016,2, 1–11. [CrossRef]
4.
Gupta, G.; Dhar, S.; Dass, A.; Sharma, V.K.; Shukla, L.; Singh, R.; Kumar, A.; Kumar, A.; Jinger, D.; Kumar, D.; et al. Assessment of
Bio-Inoculants-Mediated Nutrient Management in Terms of Productivity, Profitability and Nutrient Harvest Index of Pigeon
Pea–Wheat Cropping System in India. J. Plant Nutr. 2020,43, 2911–2928. [CrossRef]
5.
Timofeeva, A.M.; Galyamova, M.R.; Sedykh, S.E. Plant Growth-Promoting Bacteria of Soil: Designing of Consortia Beneficial for
Crop Production. Microorganisms 2023,11, 2864. [CrossRef]
6.
Mahmud, K.; Makaju, S.; Ibrahim, R.; Missaoui, A. Current Progress in Nitrogen Fixing Plants and Microbiome Research. Plants
2020,9, 97. [CrossRef]
7.
Timofeeva, A.M.; Galyamova, M.R.; Sedykh, S.E. Bacterial Siderophores: Classification, Biosynthesis, Perspectives of Use in
Agriculture. Plants 2022,11, 3065. [CrossRef]
8.
Timofeeva, A.; Galyamova, M.; Sedykh, S. Prospects for Using Phosphate-Solubilizing Microorganisms as Natural Fertilizers in
Agriculture. Plants 2022,11, 2119. [CrossRef]
9.
Castaldi, S.; Petrillo, C.; Donadio, G.; Piaz, F.D.; Cimmino, A.; Masi, M.; Evidente, A.; Isticato, R. Plant Growth Promotion
Function of Bacillus sp. Strains Isolated from Salt-Pan Rhizosphere and Their Biocontrol Potential against Macrophomina phaseolina.
Int. J. Mol. Sci. 2021,22, 3324. [CrossRef]
10.
Petrillo, C.; Castaldi, S.; Lanzilli, M.; Selci, M.; Cordone, A.; Giovannelli, D.; Isticato, R. Genomic and Physiological Characteriza-
tion of Bacilli Isolated from Salt-Pans with Plant Growth Promoting Features. Front. Microbiol. 2021,12, 715678. [CrossRef]
11.
Kour, D.; Rana, K.L.; Yadav, A.N.; Sheikh, I.; Kumar, V.; Dhaliwal, H.S.; Saxena, A.K. Amelioration of Drought Stress in Foxtail
Millet (Setaria italica L.) by P-Solubilizing Drought-Tolerant Microbes with Multifarious Plant Growth Promoting Attributes.
Environ. Sustain. 2020,3, 23–34. [CrossRef]
12.
Timmusk, S.; Behers, L.; Muthoni, J.; Muraya, A.; Aronsson, A.-C. Perspectives and Challenges of Microbial Application for Crop
Improvement. Front. Plant Sci. 2017,8, 49. [CrossRef] [PubMed]
13.
Lazali, M.; Bargaz, A. Examples of Belowground Mechanisms Enabling Legumes to Mitigate Phosphorus Deficiency. In Legume
Nitrogen Fixation in Soils with Low Phosphorus Availability; Springer International Publishing: Cham, Switzerland, 2017; pp. 135–152.
14.
Roughley, R.J.; Gault, R.R.; Gemell, L.G.; Andrews, J.A.; Brockwell, J.; Dunn, B.W.; Griffiths, G.W.; Hartley, E.J.; Hebb, D.M.;
Peoples, M.B.; et al. Autecology of Bradyrhizobium Japonicum in Soybean-Rice Rotations. Plant Soil
1995
,176, 7–14. [CrossRef]
15.
Ladha, J.K.; Tirol-Padre, A.; Reddy, C.K.; Cassman, K.G.; Verma, S.; Powlson, D.S.; van Kessel, C.; Richter, D.d.B.; Chakraborty, D.;
Pathak, H. Global Nitrogen Budgets in Cereals: A 50-Year Assessment for Maize, Rice and Wheat Production Systems. Sci. Rep.
2016,6, 19355. [CrossRef]
16.
Gupta, V.V.S.R.; Roper, M.M.; Roget, D.K. Potential for non-symbiotic N2-fixation in different agroecological zones of southern
Australia. Soil Res. 2006,44, 343. [CrossRef]
17.
Reed, S.C.; Cleveland, C.C.; Townsend, A.R. Functional Ecology of Free-Living Nitrogen Fixation: A Contemporary Perspective.
Annu. Rev. Ecol. Evol. Syst. 2011,42, 489–512. [CrossRef]
Plants 2023,12, 4074 12 of 16
18.
Malik, A.I.; Colmer, T.D.; Lambers, H.; Setter, T.L.; Schortemeyer, M. Short-Term Waterlogging has Long-Term Effects on the
Growth and Physiology of Wheat. New Phytol. 2002,153, 225–236. [CrossRef]
19.
Ritika, B.; Utpal, D. Biofertilizer, a Way Towards Organic Agriculture: A Review. Afr. J. Microbiol. Res.
2014
,8, 2332–2343.
[CrossRef]
20.
Barron, A.R.; Wurzburger, N.; Bellenger, J.P.; Wright, S.J.; Kraepiel, A.M.L.; Hedin, L.O. Molybdenum Limitation of Asymbiotic
Nitrogen Fixation in Tropical Forest Soils. Nat. Geosci. 2009,2, 42–45. [CrossRef]
21.
Ludden, P.W. Reversible ADP-Ribosylation as a Mechanism of Enzyme Regulation in Procaryotes. Mol. Cell. Biochem.
1994
,138,
123–129. [CrossRef]
22.
Robson, R.L.; Postgate, J.R. Oxygen and Hydrogen in Biological Nitrogen Fixation. Annu. Rev. Microbiol.
1980
,34, 183–207.
[CrossRef] [PubMed]
23.
Ladha, J.K.; Peoples, M.B.; Reddy, P.M.; Biswas, J.C.; Bennett, A.; Jat, M.L.; Krupnik, T.J. Biological Nitrogen Fixation and Prospects
for Ecological Intensification in Cereal-Based Cropping Systems. Field Crop. Res. 2022,283, 108541. [CrossRef] [PubMed]
24.
Eady, R.R. Structure
Function Relationships of Alternative Nitrogenases. Chem. Rev.
1996
,96, 3013–3030. [CrossRef] [PubMed]
25.
Raymond, J.; Siefert, J.L.; Staples, C.R.; Blankenship, R.E. The Natural History of Nitrogen Fixation. Mol. Biol. Evol.
2004
,21,
541–554. [CrossRef] [PubMed]
26.
Rodríguez, H.; Fraga, R. Phosphate Solubilizing Bacteria and their Role in Plant Growth Promotion. Biotechnol. Adv.
1999
,17,
319–339. [CrossRef] [PubMed]
27.
Delavaux, C.S.; Smith-Ramesh, L.M.; Kuebbing, S.E. Beyond Nutrients: A Meta-Analysis of the Diverse Effects of Arbuscular
Mycorrhizal Fungi on Plants and Soils. Ecology 2017,98, 2111–2119. [CrossRef] [PubMed]
28.
Zhu, F.; Qu, L.; Hong, X.; Sun, X. Isolation and Characterization of a Phosphate-Solubilizing Halophilic BacteriumKushneriasp.
YCWA18 from Daqiao Saltern on the Coast of Yellow Sea of China. Evid.-Based Complement. Altern. Med.
2011
,2011, 615032.
[CrossRef]
29.
Sharma, S.B.; Sayyed, R.Z.; Trivedi, M.H.; Gobi, T.A. Phosphate solubilizing microbes: Sustainable approach for managing
phosphorus deficiency in agricultural soils. Springer Plus 2013,2, 587. [CrossRef]
30.
Kalayu, G. Phosphate Solubilizing Microorganisms: Promising Approach as Biofertilizers. Int. J. Agron.
2019
,2019, 4917256.
[CrossRef]
31.
Arai, Y.; Sparks, D.L. Phosphate Reaction Dynamics in Soils and Soil Components: A Multiscale Approach. Adv. Agron.
2007
,94,
135–179.
32.
Seshachala, U.; Tallapragada, P. Phosphate Solubilizers from the Rhizospher of Piper nigrum L. in Karnataka, India. Chil. J. Agric.
Res. 2012,72, 397–403. [CrossRef]
33.
Zeng, Q.; Wu, X.; Wang, J.; Ding, X. Phosphate Solubilization and Gene Expression of Phosphate-Solubilizing Bacterium
Burkholderia multivorans WS-FJ9 under Different Levels of Soluble Phosphate. J. Microbiol. Biotechnol.
2017
,27, 844–855.
[CrossRef] [PubMed]
34.
Oubrie, A. Structure and Mechanism of Soluble Quinoprotein Glucose Dehydrogenase. EMBO J.
1999
,18, 5187–5194. [CrossRef]
[PubMed]
35.
Lessie, T.G.; Phibbs, P.V. Alternative Pathways of Carbohydrate Utilization in Pseudomonads. Annu. Rev. Microbiol.
1984
,38,
359–388. [CrossRef] [PubMed]
36.
Peix, A.; Mateos, P.F.; Rodriguez-Barrueco, C.; Martinez-Molina, E.; Velazquez, E. Growth promotion of common bean (Phaseolus
vulgaris L.) by a strain of Burkholderia cepacia under growth chamber conditions. Soil Biol. Biochem.
2001
,33, 1927–1935. [CrossRef]
37.
Nannipieri, P.; Giagnoni, L.; Landi, L.; Renella, G. Role of Phosphatase Enzymes in Soil. In Phosphorus in Action; Springer:
Berlin/Heidelberg, Germany, 2010; pp. 215–243. [CrossRef]
38.
Jorquera, M.A.; Crowley, D.E.; Marschner, P.; Greiner, R.; Fernández, M.T.; Romero, D.; Menezes-Blackburn, D.; Mora, M.D.L.L.
Identification of
β
-Propeller Phytase-Encoding Genes in Culturable paenibacillus and Bacillus spp. from the Rhizosphere of Pasture
Plants on Volcanic Soils. FEMS Microbiol. Ecol. 2011,75, 163–172. [CrossRef]
39.
Richardson, A.E.; Simpson, R.J. Soil Microorganisms Mediating Phosphorus Availability Update on Microbial Phosphorus. Plant
Physiol. 2011,156, 989–996. [CrossRef]
40.
Lidbury, I.D.E.A.; Murphy, A.R.J.; Scanlan, D.J.; Bending, G.D.; Jones, A.M.E.; Moore, J.D.; Goodall, A.; Hammond, J.P.; Wellington,
E.M.H. Comparative genomic, proteomic and exoproteomic analyses of three Pseudomonas strains reveals novel insights into the
phosphorus scavenging capabilities of soil bacteria. Environ. Microbiol. 2016,18, 3535–3549. [CrossRef]
41.
Jorquera, M.A.; Hernández, M.T.; Rengel, Z.; Marschner, P.; de la Luz Mora, M. Isolation of Culturable Phosphobacteria with Both
Phytate-Mineralization and Phosphate-Solubilization Activity from the Rhizosphere of Plants Grown in a Volcanic Soil. Biol.
Fertil. Soils 2008,44, 1025–1034. [CrossRef]
42. Rodríguez, H.; Rossolini, G.M.; Gonzalez, T.; Li, J.P.; Glick, B.R. Isolation of a Gene from Burkholderia cepacia IS-16 Encoding a
Protein That Facilitates Phosphatase Activity. Curr. Microbiol. 2000,40, 362–366. [CrossRef]
43.
Thaller, M.C.; Berlutti, F.; Schippa, S.; Iori, P.; Passariello, C.; Rossolini, G.M. Heterogeneous Patterns of Acid Phosphatases
Containing Low-Molecular-Mass Polypeptides in Members of the Family Enterobacteriaceae. Int. J. Syst. Microbiol.
1995
,45,
255–261. [CrossRef]
44.
Ohtake, H.; Wu, H.; Imazu, K.; Anbe, Y.; Kato, J.; Kuroda, A. Bacterial Phosphonate Degradation, Phosphite Oxidation and
Polyphosphate Accumulation. Resour. Conserv. Recycl. 1996,18, 125–134. [CrossRef]
Plants 2023,12, 4074 13 of 16
45.
Richardson, A.E.; Hadobas, P.A. Soil Isolates of Pseudomonas spp. that Utilize Inositol Phosphates. Can. J. Microbiol.
1997
,43,
509–516. [CrossRef] [PubMed]
46.
Heuck, C.; Smolka, G.; Whalen, E.D.; Frey, S.; Gundersen, P.; Moldan, F.; Fernandez, I.J.; Spohn, M. Effects of Long-Term Nitrogen
Addition on Phosphorus Cycling in Organic Soil Horizons of Temperate Forests. Biogeochemistry 2018,141, 167–181. [CrossRef]
47.
Marklein, A.R.; Houlton, B.Z. Nitrogen Inputs Accelerate Phosphorus Cycling Rates Across a Wide Variety of Terrestrial
Ecosystems. New Phytol. 2012,193, 696–704. [CrossRef]
48. Ganz, T.; Nemeth, E. Iron Homeostasis in Host Defence and Inflammation. Nat. Rev. Immunol. 2015,15, 500–510. [CrossRef]
49.
Miethke, M.; Marahiel, M.A. Siderophore-Based Iron Acquisition and Pathogen Control. Microbiol. Mol. Biol. Rev.
2007
,71,
413–451. [CrossRef]
50.
Raymond, K.N.; Dertz, E.A.; Kim, S.S. Enterobactin: An Archetype for Microbial Iron Transport. Proc. Natl. Acad. Sci. USA
2003
,
100, 3584–3588. [CrossRef]
51.
Li, K.; Chen, W.-H.; Bruner, S.D. Microbial Siderophore-Based Iron Assimilation and Therapeutic Applications. BioMetals
2016
,29,
377–388. [CrossRef]
52. Ratledge, C.; Dover, L.G. Iron Metabolism in Pathogenic Bacteria. Annu. Rev. Microbiol. 2000,54, 881–941. [CrossRef]
53.
Dertz, E.A.; Xu, J.; Stintzi, A.; Raymond, K.N. Bacillibactin-Mediated Iron Transport in Bacillus subtilis.J. Am. Chem. Soc.
2006
,128,
22–23. [CrossRef] [PubMed]
54.
Scavino, A.F.; Pedraza, R.O. The Role of Siderophores in Plant Growth-Promoting Bacteria. In Bacteria in Agrobiology: Crop
Productivity; Springer: Berlin/Heidelberg, Germany, 2013; pp. 265–285.
55.
Khasheii, B.; Mahmoodi, P.; Mohammadzadeh, A. Siderophores: Importance in bacterial pathogenesis and applications in
medicine and industry. Microbiol. Res. 2021,250, 126790. [CrossRef] [PubMed]
56.
Renshaw, J.C.; Robson, G.D.; Trinci, A.P.J.; Wiebe, M.G.; Livens, F.R.; Collison, D.; Taylor, R.J. Fungal siderophores: Structures,
functions and applications. Mycol. Res. 2002,106, 1123–1142. [CrossRef]
57.
Pecoraro, L.; Wang, X.; Shah, D.; Song, X.; Kumar, V.; Shakoor, A.; Tripathi, K.; Ramteke, P.W.; Rani, R. Biosynthesis Pathways,
Transport Mechanisms and Biotechnological Applications of Fungal Siderophores. J. Fungi 2021,8, 21. [CrossRef]
58.
Zabrieski, Z.; Morrell, E.; Hortin, J.; Dimkpa, C.; McLean, J.; Britt, D.; Anderson, A. Pesticidal Activity of Metal Oxide Nanoparti-
cles on Plant Pathogenic Isolates of Pythium. Ecotoxicology 2015,24, 1305–1314. [CrossRef]
59. Hider, R.C.; Kong, X. Chemistry and Biology of Siderophores. Nat. Prod. Rep. 2010,27, 637–657. [CrossRef] [PubMed]
60.
Barry, S.M.; Challis, G.L. Recent Advances in Siderophore Biosynthesis. Curr. Opin. Chem. Biol.
2009
,13, 205–215. [CrossRef]
[PubMed]
61.
Ustiatik, R.; Nuraini, Y.; Suharjono, S.; Handayanto, E. Siderophore Production of the Hg-Resistant Endophytic Bacteria Isolated
from Local Grass in the Hg-Contaminated Soil. J. Ecol. Eng. 2021,22, 129–138. [CrossRef]
62.
Butler, A.; Theisen, R.M. Iron(III)–Siderophore Coordination Chemistry: Reactivity of Marine Siderophores. Coord. Chem. Rev.
2010,254, 288–296. [CrossRef]
63.
Byers, B.R.; Powell, M.V.; Lankford, C.E. Iron-chelating Hydroxamic Acid (Schizokinen) Active in Initiation of Cell Division in
Bacillus megaterium.J. Bacteriol. 1967,93, 286–294. [CrossRef]
64.
Storey, E.P.; Boghozian, R.; Little, J.L.; Lowman, D.W.; Chakraborty, R. Characterization of ‘Schizokinen’; a Dihydroxamate-type
Siderophore Produced by Rhizobium Leguminosarum IARI 917. BioMetals 2006,19, 637–649. [CrossRef] [PubMed]
65.
Kamber, T.; Lansdell, T.A.; Stockwell, V.O.; Ishimaru, C.A.; Smits, T.H.M.; Duffy, B. Characterization of the Biosynthetic Operon
for the Antibacterial Peptide Herbicolin in Pantoea vagans Biocontrol Strain C9-1 and Incidence in Pantoea Species. Appl. Environ.
Microbiol. 2012,78, 4412–4419. [CrossRef] [PubMed]
66.
Smith, M.J.; Shoolery, J.N.; Schwyn, B.; Holden, I.; Neilands, J.B. Rhizobactin, a Structurally Novel Siderophore from Rhizobium
Meliloti. J. Am. Chem. Soc. 1985,107, 1739–1743. [CrossRef]
67. Wright, W.; Little, J.; Liu, F.; Chakraborty, R. Isolation and structural identification of the trihydroxamate siderophore vicibactin
and its degradative products from Rhizobium leguminosarum ATCC 14479 bv. trifolii.BioMetals 2013,26, 271–283. [CrossRef]
68.
Aguirre-Noyola, J.L.; Rosenblueth, M.; Santiago-Martínez, M.G.; Martínez-Romero, E. Transcriptomic Responses of Rhizobium
phaseoli to Root Exudates Reflect Its Capacity to Colonize Maize and Common Bean in an Intercropping System. Front. Microbiol.
2021,12, 740818. [CrossRef]
69.
Ghosh, S.K.; Bera, T.; Chakrabarty, A.M. Microbial Siderophore—A Boon to Agricultural Sciences. Biol. Control
2020
,144, 104214.
[CrossRef]
70.
Mishra, A.; Baek, K.-H. Salicylic Acid Biosynthesis and Metabolism: A Divergent Pathway for Plants and Bacteria. Biomolecules
2021,11, 705. [CrossRef]
71.
Bachhawat, A.K.; Ghosh, S. Iron Transport in Azospirillum brasilense: Role of the Siderophore Spirilobactin. Microbiology
1987
,
133, 1759–1765. [CrossRef]
72.
Shah, S.; Karkhanis, V.; Desai, A. Isolation and Characterization of Siderophore, with Antimicrobial Activity, from Azospirillum
Lipoferum M. Curr. Microbiol. 1992,25, 347–351. [CrossRef]
73.
Tindale, A.E.; Mehrotra, M.; Ottem, D.; Page, W.J. Dual Regulation of Catecholate Siderophore Biosynthesis in Azotobacter
Vinelandii by Iron and Oxidative Stress The GenBank Accession Number for the Sequence Reported in this Paper is AF238500.
Microbiology 2000,146, 1617–1626. [CrossRef]
Plants 2023,12, 4074 14 of 16
74.
Kraepiel, A.M.L.; Bellenger, J.P.; Wichard, T.; Morel, F.M.M. Multiple Roles of siderophores in Free-Living Nitrogen-Fixing
Bacteria. BioMetals 2009,22, 573–581. [CrossRef]
75.
Eng-Wilmot, D.L.; Van der Helm, D. Molecular and Crystal Structure of the Linear Tricatechol Siderophore, Agrobactin. J. Am.
Chem. Soc. 1980,102, 7719–7725. [CrossRef]
76.
Ong, S.A.; Peterson, T.; Neilands, J.B. Agrobactin, a Siderophore from Agrobacterium Tumefaciens. J. Biol. Chem.
1979
,254,
1860–1865. [CrossRef] [PubMed]
77.
Ito, T. Enzymatic Determination of Itoic Acid, a Bacillus Subtilis Siderophore, and 2,3-Dihydroxybenzoic Acid. Appl. Environ.
Microbiol. 1993,59, 2343–2345. [CrossRef]
78.
Wilson, M.K.; Abergel, R.J.; Raymond, K.N.; Arceneaux, J.E.L.; Byers, B.R. Siderophores of Bacillus anthracis,Bacillus cereus, and
Bacillus thuringiensis.Biochem. Biophys. Res. Commun. 2006,348, 320–325. [CrossRef]
79.
Budzikiewicz, H. Secondary Metabolites from Fluorescent pseudomonads.FEMS Microbiol. Lett.
1993
,104, 209–228. [CrossRef]
[PubMed]
80. Budzikiewicz, H. Siderophores of Fluorescent pseudomonads.Z. Für Naturforsch. C 1997,52, 713–720. [CrossRef]
81. Ringel, M.T.; Brüser, T. The Biosynthesis of Pyoverdines. Microb. Cell 2018,5, 424–437. [CrossRef] [PubMed]
82.
Philson, S.B.; Llinás, M. Siderochromes from Pseudomonas fluorescens. I. Isolation and characterization. J. Biol. Chem.
1982
,257,
8081–8085. [CrossRef]
83.
Youard, Z.A.; Mislin, G.L.A.; Majcherczyk, P.A.; Schalk, I.J.; Reimmann, C. Pseudomonas fluorescens CHA0 Produces Enantio-
pyochelin, the Optical Antipode of the Pseudomonas aeruginosa Siderophore Pyochelin. J. Biol. Chem.
2007
,282, 35546–35553.
[CrossRef]
84.
Mossialos, D.; Meyer, J.-M.; Budzikiewicz, H.; Wolff, U.; Koedam, N.; Baysse, C.; Anjaiah, V.; Cornelis, P. Quinolobactin, a New
Siderophore of Pseudomonas fluorescens ATCC 17400, the Production of Which Is Repressed by the Cognate Pyoverdine. Appl.
Environ. Microbiol. 2000,66, 487–492. [CrossRef] [PubMed]
85.
Matthijs, S.; Budzikiewicz, H.; Schäfer, M.; Wathelet, B.; Cornelis, P. Ornicorrugatin, a New Siderophore from Pseudomonas
fluorescens AF76. Z. Für Naturforsch. C 2008,63, 8–12. [CrossRef] [PubMed]
86.
Campestre, M.P.; Castagno, L.N.; Estrella, M.J.; Ruiz, O.A. Lotus Japonicus Plants of the Gifu B-129 Ecotype Subjected to Alkaline
Stress Improve their Fe
2+
bio-Availability through Inoculation with Pantoea Eucalypti M91. J. Plant Physiol.
2016
,192, 47–55.
[CrossRef] [PubMed]
87.
Bulen, W.A.; LeComte, J.R. Isolation and Properties of a Yellow-Green Fluorescent Peptide from Azotobacter Medium. Biochem.
Biophys. Res. Commun. 1962,9, 523–528. [CrossRef] [PubMed]
88.
Sharma, A.; Johri, B.N. Growth Promoting Influence of Siderophore-Producing Pseudomonas Strains GRP3A and PRS9 in Maize
(Zea mays L.) under Iron Limiting Conditions. Microbiol. Res. 2003,158, 243–248. [CrossRef]
89.
Ambrosi, C.; Leoni, L.; Putignani, L.; Orsi, N.; Visca, P. Pseudobactin Biogenesis in the Plant Growth-Promoting Rhizobacterium
Pseudomonas Strain B10: Identification and Functional Analysis of the l -Ornithine N
5
-Oxygenase (psbA) Gene. J. Bacteriol.
2000
,
182, 6233–6238. [CrossRef]
90.
Gründlinger, M.; Yasmin, S.; Lechner, B.E.; Geley, S.; Schrettl, M.; Hynes, M.; Haas, H. Fungal Siderophore Biosynthesis is Partially
Localized in Peroxisomes. Mol. Microbiol. 2013,88, 862–875. [CrossRef]
91.
Ahmed, E.; Holmström, S.J.M. Siderophores in environmental research: Roles and applications. Microb. Biotechnol.
2014
,7,
196–208. [CrossRef]
92. Dick, R. Soil Health: The Theory of Everything (Terrestrial) or Just Another Buzzword? CSA News 2019,63, 12–17. [CrossRef]
93.
Valenzuela-Aragon, B.; Parra-Cota, F.I.; Santoyo, G.; Arellano-Wattenbarger, G.L.; de los Santos-Villalobos, S. Plant-Assisted
Selection: A Promising Alternative for
in vivo
Identification of Wheat (Triticum turgidum L. subsp. Durum) Growth Promoting
Bacteria. Plant Soil 2018,435, 367–384. [CrossRef]
94.
Zuluaga, M.Y.A.; Lima Milani, K.M.; Azeredo Gonçalves, L.S.; Martinez de Oliveira, A.L. Diversity and Plant Growth-Promoting
Functions of Diazotrophic/N-Scavenging Bacteria Isolated from the Soils and Rhizospheres of Two Species of Solanum. PLoS
ONE 2020,15, e0227422. [CrossRef] [PubMed]
95.
Ortega-Urquieta, M.E.; Valenzuela-Ruíz, V.; Mitra, D.; Hyder, S.; Elsheery, N.I.; Das Mohapatra, P.K.; Parra-Cota, F.I.; Santos-de
los Villalobos, S. Draft Genome Sequence of Priestia sp. Strain TSO9, a Plant Growth-Promoting Bacterium Associated with Wheat
(Triticum turgidum subsp. durum) in the Yaqui Valley, Mexico. Plants 2022,11, 2231. [CrossRef] [PubMed]
96.
Cheng, H.-R.; Jiang, N. Extremely Rapid Extraction of DNA from Bacteria and Yeasts. Biotechnol. Lett.
2006
,28, 55–59. [CrossRef]
[PubMed]
97.
Hu, X.; Chen, J.; Guo, J. Two Phosphate- and Potassium-solubilizing Bacteria Isolated from Tianmu Mountain, Zhejiang, China.
World J. Microbiol. Biotechnol. 2006,22, 983–990. [CrossRef]
98.
Fiodor, A.; Ajijah, N.; Dziewit, L.; Pranaw, K. Biopriming of Seed with Plant Growth-Promoting Bacteria for Improved Germination
and Seedling Growth. Front. Microbiol. 2023,14, 1142966. [CrossRef]
99.
Boubekri, K.; Soumare, A.; Mardad, I.; Lyamlouli, K.; Hafidi, M.; Ouhdouch, Y.; Kouisni, L. The Screening of Potassium- and
Phosphate-Solubilizing Actinobacteria and the Assessment of Their Ability to Promote Wheat Growth Parameters. Microorganisms
2021,9, 470. [CrossRef]
100. Schwyn, B.; Neilands, J.B. Universal Chemical Assay for the Detection and Determination of Siderophores. Anal. Biochem. 1987,
160, 47–56. [CrossRef] [PubMed]
Plants 2023,12, 4074 15 of 16
101. Patel, N. Siderophores; Springer: New York, NY, USA, 2022; pp. 351–359; ISBN 978-1-0716-1724-3.
102.
Arnow, L.E. Proposed Chemical Mechanisms for The Production of Skin Erythema and Pigmentation By Radiant Energy. Science
1937,86, 176. [CrossRef]
103.
Csáky, T.Z.; Hassel, O.; Rosenberg, T.; Lång (Loukamo), S.; Turunen, E.; Tuhkanen, A. On the Estimation of Bound Hydroxylamine
in Biological Materials. Acta Chem. Scand. 1948,2, 450–454. [CrossRef]
104.
Shenker, M.; Oliver, I.; Helmann, M.; Hadar, Y.; Chen, Y. Utilization by Tomatoes of Iron Mediated by a Siderophore Produced by
Rhizopus arrhizus.J. Plant Nutr. 1992,15, 2173–2182. [CrossRef]
105.
Kudoyarova, G.R.; Vysotskaya, L.B.; Arkhipova, T.N.; Kuzmina, L.Y.; Galimsyanova, N.F.; Sidorova, L.V.; Gabbasova, I.M.;
Melentiev, A.I.; Veselov, S.Y. Effect of Auxin Producing and Phosphate Solubilizing Bacteria on Mobility of Soil Phosphorus,
Growth Rate, and P Acquisition by Wheat Plants. Acta Physiol. Plant. 2017,39, 253. [CrossRef]
106.
Saraf, M.; Jha, C.K.; Patel, D. The Role of ACC Deaminase Producing PGPR in Sustainable Agriculture. In Plant Growth and Health
Promoting Bacteria; Springer: Berlin/Heidelberg, Germany, 2010; pp. 365–385.
107.
Baudoin, E.; Lerner, A.; Mirza, M.S.; El Zemrany, H.; Prigent-Combaret, C.; Jurkevich, E.; Spaepen, S.; Vanderleyden, J.; Nazaret,
S.; Okon, Y.; et al. Effects of Azospirillum brasilense with genetically modified auxin biosynthesis gene ipdC upon the diversity
of the indigenous microbiota of the wheat rhizosphere. Res. Microbiol. 2010,161, 219–226. [CrossRef] [PubMed]
108. Yue, J.; Hu, X.; Huang, J. Origin of plant auxin biosynthesis. Trends Plant Sci. 2014,19, 764–770. [CrossRef] [PubMed]
109.
Spaepen, S.; Vanderleyden, J.; Remans, R. Indole-3-Acetic Acid in Microbial and Microorganism-Plant Signaling. FEMS Microbiol.
Rev. 2007,31, 425–448. [CrossRef] [PubMed]
110.
Espindula, E.; Sperb, E.R.; Moz, B.; Pankievicz, V.C.S.; Tuleski, T.R.; Tadra-Sfeir, M.Z.; Bonato, P.; Scheid, C.; Merib, J.; de Souza,
E.M.; et al. Effects on gene expression during maize-Azospirillum interaction in the presence of a plant-specific inhibitor of
indole-3-acetic acid production. Genet. Mol. Biol. 2023,46, e20230100. [CrossRef]
111. Gordon, S.A.; Weber, R.P. Colorimetric Estimation of Indoleacetic Acid. Plant Physiol. 1951,26, 192–195. [CrossRef]
112.
Glickmann, E.; Dessaux, Y. A Critical Examination of the Specificity of the Salkowski Reagent for Indolic Compounds Produced
by Phytopathogenic Bacteria. Appl. Environ. Microbiol. 1995,61, 793–796. [CrossRef]
113.
Santiago, C.D.; Yagi, S.; Ijima, M.; Nashimoto, T.; Sawada, M.; Ikeda, S.; Asano, K.; Orikasa, Y.; Ohwada, T. Bacterial Compatibility
in Combined Inoculations Enhances the Growth of Potato Seedlings. Microbes Environ. 2017,32, 14–23. [CrossRef]
114.
Bal, H.B.; Nayak, L.; Das, S.; Adhya, T.K. Isolation of ACC Deaminase Producing PGPR from Rice Rhizosphere and Evaluating
their Plant Growth Promoting Activity under Salt Stress. Plant Soil 2013,366, 93–105. [CrossRef]
115.
Pranaw, K.; Pidlisnyuk, V.; Trögl, J.; Malinská, H. Bioprospecting of a Novel Plant Growth-Promoting Bacterium Bacillus Altitudinis
KP-14 for Enhancing Miscanthus ×giganteus Growth in Metals Contaminated Soil. Biology 2020,9, 305. [CrossRef]
116.
Hernández-León, R.; Rojas-Solís, D.; Contreras-Pérez, M.; Orozco-Mosqueda, M.d.C.; Macías-Rodríguez, L.I.; Reyes-de la Cruz,
H.; Valencia-Cantero, E.; Santoyo, G. Characterization of the Antifungal and Plant Growth-Promoting Effects of Diffusible and
Volatile Organic Compounds Produced by Pseudomonas Fluorescens Strains. Biol. Control 2015,81, 83–92. [CrossRef]
117.
López-Bucio, J.; Hernández-Abreu, E.; Sánchez-Calderón, L.; Nieto-Jacobo, M.F.; Simpson, J.; Herrera-Estrella, L. Phosphate
Availability Alters Architecture and Causes Changes in Hormone Sensitivity in the Arabidopsis Root System. Plant Physiol.
2002
,
129, 244–256. [CrossRef] [PubMed]
118.
Wisniewski-Dyé, F.; Lozano, L.; Acosta-Cruz, E.; Borland, S.; Drogue, B.; Prigent-Combaret, C.; Rouy, Z.; Barbe, V.; Herrera, A.M.;
González, V.; et al. Genome Sequence of Azospirillum brasilense CBG497 and Comparative Analyses of Azospirillum Core and
Accessory Genomes provide Insight into Niche Adaptation. Genes 2012,3, 576–602. [CrossRef] [PubMed]
119.
Alori, E.T.; Glick, B.R.; Babalola, O.O. Microbial Phosphorus Solubilization and Its Potential for Use in Sustainable Agriculture.
Front. Microbiol. 2017,8, 971. [CrossRef] [PubMed]
120.
Haidar, B.; Ferdous, M.; Fatema, B.; Ferdous, A.S.; Islam, M.R.; Khan, H. Population Diversity of Bacterial Endophytes from Jute
(Corchorus olitorius) and Evaluation of Their Potential Role as Bioinoculants. Microbiol. Res.
2018
,208, 43–53. [CrossRef] [PubMed]
121.
Thomas, L.; Hodgson, D.A.; Wentzel, A.; Nieselt, K.; Ellingsen, T.E.; Moore, J.; Morrissey, E.R.; Legaie, R.; The STREAM
Consortium; Wohlleben, W.; et al. Metabolic Switches and Adaptations Deduced from the Proteomes of Streptomyces coelicolor
Wild Type and phoP Mutant Grown in Batch Culture. Mol. Cell. Proteom. 2012,11, M111.013797. [CrossRef] [PubMed]
122.
Buch, A.; Archana, G.; Naresh Kumar, G. Metabolic Channeling of Glucose Towards Gluconate in Phosphate-Solubilizing
Pseudomonas Aeruginosa P4 under Phosphorus Deficiency. Res. Microbiol. 2008,159, 635–642. [CrossRef]
123.
Zeng, Q.; Wu, X.; Wen, X. Effects of Soluble Phosphate on Phosphate-Solubilizing Characteristics and Expression of gcd Gene in
Pseudomonas frederiksbergensis JW-SD2. Curr. Microbiol. 2016,72, 198–206. [CrossRef]
124.
Ramakrishna, W.; Rathore, P.; Kumari, R.; Yadav, R. Brown Gold of Marginal Soil: Plant Growth Promoting Bacteria to Overcome
Plant Abiotic Stress for Agriculture, Biofuels and Carbon Sequestration. Sci. Total. Environ. 2020,711, 135062. [CrossRef]
125.
Hu, J.; Wei, Z.; Weidner, S.; Friman, V.-P.; Xu, Y.-C.; Shen, Q.-R.; Jousset, A. Probiotic Pseudomonas Communities Enhance
Plant Growth and Nutrient Assimilation via Diversity-Mediated Ecosystem Functioning. Soil Biol. Biochem.
2017
,113, 122–129.
[CrossRef]
126.
Assainar, S.K.; Abbott, L.K.; Mickan, B.S.; Whiteley, A.S.; Siddique, K.H.M.; Solaiman, Z.M. Response of Wheat to a Multiple
Species Microbial Inoculant Compared to Fertilizer Application. Front. Plant Sci. 2018,9, 1601. [CrossRef] [PubMed]
Plants 2023,12, 4074 16 of 16
127.
Agaras, B.C.; Scandiani, M.; Luque, A.; Fernández, L.; Farina, F.; Carmona, M.; Gally, M.; Romero, A.; Wall, L.; Valverde, C.
Quantification of the Potential Biocontrol and Direct Plant Growth Promotion Abilities based on Multiple Biological Traits
Distinguish Different Groups of Pseudomonas spp. Isolates. Biol. Control 2015,90, 173–186. [CrossRef]
128.
Wei, Z.; Yang, T.; Friman, V.-P.; Xu, Y.; Shen, Q.; Jousset, A. Trophic Network Architecture of Root-Associated Bacterial
Communities Determines Pathogen Invasion and Plant Health. Nat. Commun. 2015,6, 8413. [CrossRef] [PubMed]
129.
Hu, J.; Wei, Z.; Friman, V.-P.; Gu, S.; Wang, X.; Eisenhauer, N.; Yang, T.; Ma, J.; Shen, Q.; Xu, Y.; et al. Probiotic Diversity Enhances
Rhizosphere Microbiome Function and Plant Disease Suppression. mBio 2016,7, e01790-16. [CrossRef] [PubMed]
130.
Kadmiri, I.M.; Chaouqui, L.; Azaroual, S.E.; Sijilmassi, B.; Yaakoubi, K.; Wahby, I. Phosphate-Solubilizing and Auxin-Producing
Rhizobacteria Promote Plant Growth under Saline Conditions. Arab. J. Sci. Eng. 2018,43, 3403–3415. [CrossRef]
131.
Chatterjee, S.; Sau, G.B.; Sinha, S.; Mukherjee, S.K. Effect of Co-Inoculation of Plant Growth-Promoting Rhizobacteria on the
Growth of Amaranth Plants. Arch. Agron. Soil Sci. 2012,58, 1387–1397. [CrossRef]
132.
Bargaz, A.; Lyamlouli, K.; Chtouki, M.; Zeroual, Y.; Dhiba, D. Soil Microbial Resources for Improving Fertilizers Efficiency in an
Integrated Plant Nutrient Management System. Front. Microbiol. 2018,9, 1606. [CrossRef] [PubMed]
133.
Gupta, M.; Bisht, S.; Singh, B.; Gulati, A.; Tewari, R. Enhanced Biomass and Steviol Glycosides in Stevia Rebaudiana Treated with
Phosphate-Solubilizing Bacteria and Rock Phosphate. Plant Growth Regul. 2011,65, 449–457. [CrossRef]
134.
Mahanta, D.; Rai, R.K.; Dhar, S.; Varghese, E.; Raja, A.; Purakayastha, T.J. Modification of Root Properties with Phosphate
Solubilizing Bacteria and Arbuscular Mycorrhiza to Reduce Rock Phosphate Application in Soybean-Wheat Cropping System.
Ecol. Eng. 2018,111, 31–43. [CrossRef]
135.
Adnan, M.; Fahad, S.; Zamin, M.; Shah, S.; Mian, I.A.; Danish, S.; Zafar-Ul-Hye, M.; Battaglia, M.L.; Naz, R.M.M.; Saeed, B.; et al.
Coupling Phosphate-Solubilizing Bacteria with Phosphorus Supplements Improve Maize Phosphorus Acquisition and Growth
under Lime Induced Salinity Stress. Plants 2020,9, 900. [CrossRef]
136.
Park, J.H.; Bolan, N.; Megharaj, M.; Naidu, R. Concomitant Rock Phosphate Dissolution and Lead Immobilization by Phosphate
Solubilizing Bacteria (Enterobacter sp.). J. Environ. Manag. 2011,92, 1115–1120. [CrossRef] [PubMed]
137.
Manzoor, M.; Abbasi, M.K.; Sultan, T. Isolation of Phosphate Solubilizing Bacteria from Maize Rhizosphere and Their Potential
for Rock Phosphate Solubilization–Mineralization and Plant Growth Promotion. Geomicrobiol. J. 2017,34, 81–95. [CrossRef]
138.
Elhaissoufi, W.; Khourchi, S.; Ibnyasser, A.; Ghoulam, C.; Rchiad, Z.; Zeroual, Y.; Lyamlouli, K.; Bargaz, A. Phosphate Solubilizing
Rhizobacteria Could Have a Stronger Influence on Wheat Root Traits and Aboveground Physiology Than Rhizosphere P
Solubilization. Front. Plant Sci. 2020,11, 979. [CrossRef]
139.
Ben Zineb, A.; Trabelsi, D.; Ayachi, I.; Barhoumi, F.; Aroca, R.; Mhamdi, R. Inoculation with Elite Strains of Phosphate-Solubilizing
Bacteria Enhances the Effectiveness of Fertilization with Rock Phosphates. Geomicrobiol. J. 2020,37, 22–30. [CrossRef]
140.
Tahir, M.; Khalid, U.; Ijaz, M.; Shah, G.M.; Naeem, M.A.; Shahid, M.; Mahmood, K.; Ahmad, N.; Kareem, F. Combined Application
of Bio-Organic Phosphate and Phosphorus Solubilizing Bacteria (Bacillus strain MWT 14) Improve the Performance of Bread
Wheat with low Fertilizer Input under an Arid Climate. Braz. J. Microbiol. 2018,49, 15–24. [CrossRef] [PubMed]
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.
... Commonly, bacteria obtain iron by the production of low-molecular mass iron chelators referred to as siderophores which have high association constants for complexing iron. Bacillus paralicheniformis, B. haynesii, B. licheniformis and Pseudomonas aeruginosa and P. fluorescens have been identified as siderophores producers, phosphate, and zinc solubilizers in rice rhizosphere which assist in the crop nutrition [68]. It has also been reported that B. mucilaginosus and Aspergillus niger can solubilize potassium, phosphorus and silicon in soil and increase rice yield [69]. ...
... In addition to antibiotic production, these microbes inhibit the growth of plant pathogens by competing for space and nutrients acquisition, production of lytic enzymes and bacteriocins [73]. The study reported that Pseudomonas aeruginosa and P. fluorescens produce different antibiotics and inhibit fungal growth [68]. In the rice rhizosphere PGPR bacteria express modulation of phytohormone signaling, production of secondary metabolites and balance of reactive oxygen species and scavengers producing ROS scavengers. ...
Article
Full-text available
The rhizosphere niche is extremely important for the overall growth and development of plants. Evidently, it is necessary to understand the complete mechanism of plant microbe interactions of the rhizosphere for sustainable and low input productivity. To meet the increasing global food demand, rice (Oryza sativa L.) agriculture seeks optimal conditions. The unique oxic-anoxic interface of rice-growing soil has invited divergent microbes with dynamic biogeochemical cycles. This review provides the systematic analysis of microbes associated with the major biogeochemical cycles with the aim to generate better management strategies of rhizospheric microbiome in the field of rice agriculture. For instance, several methanogenic and methanotrophic bacteria in the rice rhizosphere make an equilibrium for methane concentration in the environment. The carbon sequestration in paddy soil is again done through many rhizospheric microorganisms that can directly assimilate CO2 with their photoautotrophic mode of nutrition. Also the phosphate solubilizing microbes remain to be the most important keys for the PGPR activity of the paddy ecosystem. In addition, rhizospheric microbiome remain crucial in degradation and solubilization of organo-sulfur and insoluble inorganic sulfides which can be taken by the plants. Further, this review elucidates on the advantages of using metagenomic and metaproteomic approaches as an alternative of traditional approaches to understand the overall metabolic pathways operational in paddy-field. These knowledges are expected to open new possibilities for designing the balanced microbiome used as inoculum for intensive farming and will eventually lead to exert positive impacts on rice cultivation.
... Microbes and bacteria are literally everywhere, in the soil (Timofeeva et al., 2023), in water (Cabral, 2010), in animals and humans (Dekaboruah et al., 2020), in plants (Buttimer et al., 2017), in extreme environments (Rampelotto, 2013), and even in space stations (Koehle et al., 2023). Bacteria are also used as tools in many contexts, especially in health and disease (Tanniche and Behkam, 2023). ...
... Microbes and bacteria are literally everywhere, in the soil (Timofeeva et al., 2023), in water (Cabral, 2010), in animals and humans (Dekaboruah et al., 2020), in plants (Buttimer et al., 2017), in extreme environments (Rampelotto, 2013), and even in space stations (Koehle et al., 2023). Bacteria are also used as tools in many contexts, especially in health and disease (Tanniche and Behkam, 2023). They are everything, everywhere, all at once! ...
Article
Full-text available
If scientists stick to their research expertise without collaborating with other experts in different fields, it could stall the progress of their work in a world where interdisciplinary thinking and working should be second nature. Biologists are at the forefront of this trend. That is why a consortium formed by the faculty of sciences of Tunis El Manar University, Tunisia, the GetGenome Foundation and Learn and Win, decided to organise an international conference on bacteria, a perfect field for multidisciplinarity. For 3 days, from 14 to 16 December 2023, more than 200 interdisciplinary researchers and students of life sciences and more than 20 international speakers and trainers met at the faculty of sciences in Tunis, to discuss microbiology and bacteria from different horizons, from the most fundamental to the most imaginative, with a strong focus on technologies and discoveries. This Meeting Review describes the scientific event and highlights the main results of both the conferences and the practical sessions.
... The symbiotic relationship between citrus plants and their bacteriome is gaining scientific interest, particularly in the face of challenges such as rising global temperatures, nutrient scarcity, and prevalent plant diseases that compromise yield and quality Srivastava et al., 2022). Furthermore, soil microbiomes represent a rich reservoir of resources for biotechnologically relevant processes such as nitrogen fixation, phosphate solubilization, biocatalysts, secondary metabolites, and other biomolecules, etc (Cherni et al., 2019;Timofeeva et al., 2023). Understanding soil microbiomes is a critical step towards improving aspects of citriculture and agriculture as a whole (Berg, 2009;Chaparro et al., 2012). ...
Article
Full-text available
Citrus cultivation is vital to global agriculture, necessitating a comprehensive understanding of the soil microbiome’s diversity for sustainable practices. This study provides initial insights into the bacteriome in citrus crops in Santander, Colombia, employing a holistic approach combining culture-based techniques, sequencing methods, and bioinformatics analyses. The study explores organic and non-organic cultivation systems, revealing statistically significant differences in bacterial community composition between both practices. In general, the communities are dominated by members of the Actinobacteria and Proteobacteria, along with bacterial orders Gaiellales and Burkholderiales, all contributing to intricate ecological processes. Culture-based methods aided in the isolation of potential biotechnologically relevant strains. Among them, strain CP102 showed a pronounced carboxymethylcellulose (CMC) degradation capacity. Genetic analysis of the isolate resulted in the generation of the first closed genome of a member of the species Enterobacter soli and identified an unreported 109 kb plasmid. Further genomic examination revealed genes potentially associated with cellulose degradation in this species, which provides the isolate with biotechnological potential. This research significantly advances the global understanding of citrus-associated bacteriomes, shaping future agricultural practices and promoting the development of sustainable bioproducts.
... However, indirect mechanisms are associated with the production of lytic enzymes and antibiotics, dropping the soil pH and producing exopolysaccharides. Several studies have assessed the efficacy of PGPR for maintaining a sustainable agroecosystem in normal and stressful conditions (Oleńska et al., 2020;Magnabosco et al., 2023;Timofeeva et al., 2023). Numerous articles and meta-analyses have observed the beneficial impacts of NPs on soil health and agronomic productivity as well as the variables that facilitate the ameliorative role of NPs (Urra et al., 2019;Sharma et al., 2020;Verma et al., 2022a, b). ...
Article
Full-text available
Sustainable food security and safety are major concerns on a global scale, especially in developed nations. Adverse agroclimatic conditions affect the largest agricultural-producing areas, which reduces the production of crops. Achieving sustainable food safety is challenging because of several factors, such as soil flooding/waterlogging, ultraviolet (UV) rays, acidic/sodic soil, hazardous ions, low and high temperatures, and nutritional imbalances. Plant growth-promoting rhizobacteria (PGPR) are widely employed in in-vitro conditions because they are widely recognized as a more environmentally and sustainably friendly approach to increasing crop yield in contaminated and fertile soil. Conversely, the use of nanoparticles (NPs) as an amendment in the soil has recently been proposed as an economical way to enhance the texture of the soil and improving agricultural yields. Nowadays, various research experiments have combined or individually applied with the PGPR and NPs for balancing soil elements and crop yield in response to control and adverse situations, with the expectation that both additives might perform well together. According to several research findings, interactive applications significantly increase sustainable crop yields more than PGPR or NPs alone. The present review summarized the functional and mechanistic basis of the interactive role of PGPR and NPs. However, this article focused on the potential of the research direction to realize the possible interaction of PGPR and NPs at a large scale in the upcoming years.
... In the rhizosphere, bacteria stand out as one of the most abundant and varied groups of microorganisms. They are pivotal in nutrient cycling processes such as nitrogen fixation and phosphorus and potassium solubilizations, in addition to playing key roles in suppressing diseases and fostering plant growth through the synthesis of various hormones and enzymes [11,12]. Fungi are integral to this ecosystem, with symbiotic mycorrhizal fungi forming advantageous connections with plant roots to enhance water and nutrient absorption, while free-living fungi aid in the decomposition of organic matter and further nutrient cycling [13]. ...
Article
Full-text available
Plant health is necessary for food security, which is a key determinant of secure and sustainable food production systems. Deficiency of soil nutrients and invasion of plant pathogens or insects are the main destroyers of the world’s food production. Synthetic fertilizers and chemical-based pesticides are frequently employed to combat the problems. However, these have negative impacts on microbial ecosystems and ecosystem functioning. Rhizosphere microorganisms have demonstrated their potency to improve or manage plant nutrients to encourage plant growth, resulting in increased yield and quality by converting organic and inorganic substances around the rhizosphere zone into available plant nutrients. Besides regulating nutrient availability and plant growth enhancement, rhizobacteria or fungi can restrict plant pathogens that cause disease by secreting inhibitory chemicals and boosting plant immunity to combat pests or pathogens. Thus, rhizosphere microorganisms are viewed as viable, alluring economic approaches for sustainable agriculture as biofertilizers and biopesticides. This review provides an overview of the role of rhizosphere microorganisms in soil nutrients and inducing of plant defenses. Moreover, a discussion is presented surrounding the recent consequences of employing these microorganisms and a sustainable strategy towards improving fertilization effectiveness, and encouraging stronger, more pest-resistant plants.
... Their results showed that inoculation with live microorganisms played a key role. The mechanisms by which the PGPB enhanced the morphological and yield parameters of grape are hypothesized to be their PGP traits [45], which include nitrogen fixation [46], phosphorus solubilization [47], potassium dissolution, IAA, and siderophore production [48]. Unlike the study by Wei et al. [44], who used biochar as the carrier for PGPB, our study directly inoculated PGPB into organic fertilizer, using the organic fertilizer itself as the carrier. ...
Article
Full-text available
Plant growth-promoting bacteria (PGPB) have growth-promoting and disease-resisting effects and can be used as new types of plant growth promoters. This study was conducted to determine the plant growth-promoting traits of five strains and follow a 2-year field trial to evaluate their effects on grape growth and fruit quality. The five PGPB were combined with two organic fertilizers (cow dung fertilizer and distillers’ grain fertilizer) for application on grape fields; the control group only received the corresponding organic fertilizer. The five strains showed different growth promoting abilities, as indicated by their differing production of indole acetic acid (IAA) and siderophores and ability to dissolve phosphorus and potassium, fix nitrogen, and resist saline and alkali. During the field trial, vine growth and fruit quality were significantly better in the distillers’ grain fertilizer (high nutrient content) alone treatment than in the cow dung fertilizer (low nutrient content) alone treatment. However, after the two fertilizers were inoculated with the five different PGPB, only the five treatments with cow dung fertilizer inoculated with PGPB showed significant improvement. The five treatments of cow dung fertilizer inoculated with PGPB exhibited varied impacts on plant growth and fruit quality. And the promotion effects persisted significantly after two consecutive years. Among the PGPB, Bacillus velezensis 18, B. velezensis 20, and Rahnella aquatilis 5 emerged as consistently effective performers over the two-year period, demonstrating stable and commendable outcomes. These strains are recommended for prolonged application in grape cultivation to optimize growth and yield. This study provides a theoretical reference and an experimental basis for organic fertilizer inoculated with PGPB to improve grape production.
Article
Full-text available
Plant growth-promoting bacteria are commonly used in agriculture, particularly for seed inoculation. Multispecies consortia are believed to be the most promising form of these bacteria. However, designing and modeling bacterial consortia to achieve desired phenotypic outcomes in plants is challenging. This review aims to address this challenge by exploring key antimicrobial interactions. Special attention is given to approaches for developing soil plant growth-promoting bacteria consortia. Additionally, advanced omics-based methods are analyzed that allow soil microbiomes to be characterized, providing an understanding of the molecular and functional aspects of these microbial communities. A comprehensive discussion explores the utilization of bacterial preparations in biofertilizers for agricultural applications, focusing on the intricate design of synthetic bacterial consortia with these preparations. Overall, the review provides valuable insights and strategies for intentionally designing bacterial consortia to enhance plant growth and development.
Article
Full-text available
Abstract Amongst the sustainable alternatives to increase maize production is the use of plant growth-promoting bacteria (PGPB). Azospirillum brasilense is one of the most well-known PGPB being able to fix nitrogen and produce phytohormones, especially indole-3-acetic acid - IAA. This work investigated if there is any contribution of the bacterium to the plant’s IAA levels, and how it affects the plant. To inhibit plant IAA production, yucasin, an inhibitor of the TAM/YUC pathway, was applied. Plantlets’ IAA concentration was evaluated through HPLC and dual RNA-Seq was used to analyze gene expression. Statistical differences between the group treated with yucasin and the other groups showed that A. brasilense inoculation was able to prevent the phenotype caused by yucasin concerning the number of lateral roots. Genes involved in the auxin and ABA response pathways, auxin efflux transport, and the cell cycle were regulated by the presence of the bacterium, yucasin, or both. Genes involved in the response to biotic/abiotic stress, plant disease resistance, and a D-type cellulose synthase changed their expression pattern among two sets of comparisons in which A. brasilense acted as treatment. The results suggest that A. brasilense interferes with the expression of many maize genes through an IAA-independent pathway.
Article
Full-text available
Several seed priming methods can be used to improve seed germination, seedling vigor, and to overcome abiotic stress. In addition to these benefits, only the biopriming method provides the additional benefit of biotic stress management, earning it special attention. Seed biopriming is useful in almost all crops around the world and is an environmentally friendly alternative to chemical fungicides. Biopriming usually refers to use of beneficial microorganisms, in particular plant growth-promoting bacteria (PGPB) able to survive under various harsh environmental conditions. In this study, various bacterial strains were isolated from samples of different origins, i.e., rhizospheric soil, desert sand, and sea mud. Preliminary screening of 156 bacterial isolates was conducted on the basis of their potassium (K), phosphorus (P) solubilization ability, and production of plant growth hormone, i.e., indole acetic acid (IAA). The most efficient bacteria were identified by 16S rRNA gene nucleotide sequences and further examined for their ACC deaminase activity, ammonia production, and biocontrol activity (defined via chitinolytic activity, HCN, and siderophores production). Finally, carrot seed germination assay was conducted with 10 shortlisted most potent isolates. 68.6, 58.3, and 66.7% of tested bacterial isolates were capable of P, K, and Zn solubilization, respectively. Klebsiella aerogenes AF3II1 showed the highest P and K solubilization, while isolates AF4II5, AF7II3, and PC3 showed the highest IAA synthesis ability. Serratia plymuthica EDC15 and Pseudomonas putida AF1I1 showed the strongest chitinolytic and siderophore production activity, respectively. Seven isolates demonstrated strong HCN production ability. Five isolates improved carrot seed germination. Only selected isolates with plant growth-promoting properties can improve carrot germination. The results of this study demonstrate that mainly auxins are involved in seed germination. Furthermore, the data suggest that phosphate solubilization ability may play an additional role in seed germination.
Article
Full-text available
Siderophores are synthesized and secreted by many bacteria, yeasts, fungi, and plants for Fe (III) chelation. A variety of plant-growth-promoting bacteria (PGPB) colonize the rhizosphere and contribute to iron assimilation by plants. These microorganisms possess mechanisms to produce Fe ions under iron-deficient conditions. Under appropriate conditions, they synthesize and release siderophores, thereby increasing and regulating iron bioavailability. This review focuses on various bacterial strains that positively affect plant growth and development through synthesizing siderophores. Here we discuss the diverse chemical nature of siderophores produced by plant root bacteria; the life cycle of siderophores, from their biosynthesis to the Fe–siderophore complex degradation; three mechanisms of siderophore biosynthesis in bacteria; the methods for analyzing siderophores and the siderophore-producing activity of bacteria and the methods for screening the siderophore-producing activity of bacterial colonies. Further analysis of biochemical, molecular–biological, and physiological features of siderophore synthesis by bacteria and their use by plants will allow one to create effective microbiological preparations for improving soil fertility and increasing plant biomass, which is highly relevant for sustainable agriculture.
Article
Full-text available
Strain TSO9 was isolated from a commercial field of wheat (Triticum turgidum L. subsp. durum) located in the Yaqui, Valley, Mexico. Here, the genome of this strain was sequenced, obtaining a total of 5,248,515 bp; 38.0% G + C content; 1,186,514 bp N50; and 2 L50. Based on the 16S rRNA gene sequencing, strain TSO9 was affiliated with the genus Priestia. The genome annotation of Priestia sp. TSO9 contains a total of 147 RNAs, 128 tRNAs, 1 tmRNA, and 5512 coding DNA sequences (CDS) distributed into 332 subsystems, where CDS associated with agricultural purposes were identified, such as (i) virulence, disease, and defense (57 CDS) (i.e., resistance to antibiotics and toxic compounds (34 CDS), invasion and intracellular resistance (12 CDS), and bacteriocins and ribosomally synthesized antibacterial peptides (10 CDS)), (ii) iron acquisition and metabolism (36 CDS), and (iii) secondary metabolism (4 CDS), i.e., auxin biosynthesis. In addition, subsystems related to the viability of an active ingredient for agricultural bioproducts were identified, such as (i) stress response (65 CDS). These genomic traits are correlated with the metabolic background of this strain, and its positive effects on wheat growth regulation reported in this work. Thus, further investigations of Priestia sp. TSO9 are necessary to complement findings regarding its application in agroecosystems to increase wheat yield sustainably
Article
Full-text available
Phosphates are known to be essential for plant growth and development, with phosphorus compounds being involved in various physiological and biochemical reactions. Phosphates are known as one of the most important factors limiting crop yields. The problem of phosphorus deficiency in the soil has traditionally been solved by applying phosphate fertilizers. However, chemical phosphate fertilizers are considered ineffective compared to the organic fertilizers manure and compost. Therefore, increasing the bioavailability of phosphates for plants is one of the primary goals of sustainable agriculture. Phosphate-solubilizing soil microorganisms can make soil-insoluble phosphate bioavailable for plants through solubilization and mineralization. These microorganisms are currently in the focus of interest due to their advantages, such as environmental friendliness, low cost, and high biological efficiency. In this regard, the solubilization of phosphates by soil microorganisms holds strong potential in research, and inoculation of soils or crops with phosphate-solubilizing bacteria is a promising strategy to improve plant phosphate uptake. In this review, we analyze all the species of phosphate-solubilizing bacteria described in the literature to date. We discuss key mechanisms of solubilization of mineral phosphates and mineralization of organic phosphate-containing compounds: organic acids secreted by bacteria for the mobilization of insoluble inorganic phosphates, and the enzymes hydrolyzing phosphorus-containing organic compounds. We demonstrate that phosphate-solubilizing microorganisms have enormous potency as biofertilizers since they increase phosphorus bioavailability for the plant, promote sustainable agriculture, improve soil fertility, and raise crop yields. The use of phosphate-solubilizing microbes is regarded as a new frontier in increasing plant productivity.
Article
Full-text available
The demand for nitrogen (N) for crop production increased rapidly from the middle of the twentieth century and is predicted to at least double by 2050 to satisfy the on-going improvements in productivity of major food crops such as wheat, rice and maize that underpin the staple diet of most of the world’s population. The increased demand will need to be fulfilled by the two main sources of N supply – biological nitrogen (gas) (N2) fixation (BNF) and fertilizer N supplied through the Haber-Bosch processes. BNF provides many functional benefits for agroecosystems. It is a vital mechanism for replenishing the reservoirs of soil organic N and improving the availability of soil N to support crop growth while also assisting in efforts to lower negative environmental externalities than fertilizer N. In cereal-based cropping systems, legumes in symbiosis with rhizobia contribute the largest BNF input; however, diazotrophs involved in non-symbiotic associations with plants or present as free-living N2-fixers are ubiquitous and also provide an additional source of fixed N. This review presents the current knowledge of BNF by free-living, non-symbiotic and symbiotic diazotrophs in the global N cycle, examines global and regional estimates of contributions of BNF, and discusses possible strategies to enhance BNF for the prospective benefit of cereal N nutrition. We conclude by considering the challenges of introducing in planta BNF into cereals and reflect on the potential for BNF in both conventional and alternative crop management systems to encourage the ecological intensification of cereal and legume production.
Article
Full-text available
Iron (Fe) is the fourth most abundant element on earth and represents an essential nutrient for life. As a fundamental mineral element for cell growth and development, iron is available for uptake as ferric ions, which are usually oxidized into complex oxyhydroxide polymers, insoluble under aerobic conditions. In these conditions, the bioavailability of iron is dramatically reduced. As a result, microorganisms face problems of iron acquisition, especially under low concentrations of this element. However, some microbes have evolved mechanisms for obtaining ferric irons from the extracellular medium or environment by forming small molecules often regarded as siderophores. Siderophores are high affinity iron-binding molecules produced by a repertoire of proteins found in the cytoplasm of cyanobacteria, bacteria, fungi, and plants. Common groups of siderophores include hydroxamates, catecholates, carboxylates, and hydroximates. The hydroxamate siderophores are commonly synthesized by fungi. L-ornithine is a biosynthetic precursor of siderophores, which is synthesized from multimodular large enzyme complexes through non-ribosomal peptide synthetases (NRPSs), while siderophore-Fe chelators cell wall mannoproteins (FIT1, FIT2, and FIT3) help the retention of siderophores. S. cerevisiae, for example, can express these proteins in two genetically separate systems (reductive and nonreductive) in the plasma membrane. These proteins can convert Fe (III) into Fe (II) by a ferrous-specific metalloreductase enzyme complex and flavin reductases (FREs). However, regulation of the siderophore through Fur Box protein on the DNA promoter region and its activation or repression depend primarily on the Fe availability in the external medium. Siderophores are essential due to their wide range of applications in biotechnology, medicine, bioremediation of heavy metal polluted environments, biocontrol of plant pathogens, and plant growth enhancement.
Article
Full-text available
Corn and common bean have been cultivated together in Mesoamerica for thousands of years in an intercropping system called “milpa,” where the roots are intermingled, favoring the exchange of their microbiota, including symbionts such as rhizobia. In this work, we studied the genomic expression of Rhizobium phaseoli Ch24-10 (by RNA-seq) after a 2-h treatment in the presence of root exudates of maize and bean grown in monoculture and milpa system under hydroponic conditions. In bean exudates, rhizobial genes for nodulation and degradation of aromatic compounds were induced; while in maize, a response of genes for degradation of mucilage and ferulic acid was observed, as well as those for the transport of sugars, dicarboxylic acids and iron. Ch24-10 transcriptomes in milpa resembled those of beans because they both showed high expression of nodulation genes; some genes that were expressed in corn exudates were also induced by the intercropping system, especially those for the degradation of ferulic acid and pectin. Beans grown in milpa system formed nitrogen-fixing nodules similar to monocultured beans; therefore, the presence of maize did not interfere with Rhizobium –bean symbiosis. Genes for the metabolism of sugars and amino acids, flavonoid and phytoalexin tolerance, and a T3SS were expressed in both monocultures and milpa system, which reveals the adaptive capacity of rhizobia to colonize both legumes and cereals. Transcriptional fusions of the putA gene, which participates in proline metabolism, and of a gene encoding a polygalacturonase were used to validate their participation in plant–microbe interactions. We determined the enzymatic activity of carbonic anhydrase whose gene was also overexpressed in response to root exudates.
Article
Full-text available
Massive application of chemical fertilizers and pesticides has been the main strategy used to cope with the rising crop demands in the last decades. The indiscriminate use of chemicals while providing a temporary solution to food demand has led to a decrease in crop productivity and an increase in the environmental impact of modern agriculture. A sustainable alternative to the use of agrochemicals is the use of microorganisms naturally capable of enhancing plant growth and protecting crops from pests known as Plant-Growth-Promoting Bacteria (PGPB). Aim of the present study was to isolate and characterize PGPB from salt-pans sand samples with activities associated to plant fitness increase. To survive high salinity, salt-tolerant microbes produce a broad range of compounds with heterogeneous biological activities that are potentially beneficial for plant growth. A total of 20 halophilic spore-forming bacteria have been screened in vitro for phyto-beneficial traits and compared with other two members of Bacillus genus recently isolated from the rhizosphere of the same collection site and characterized as potential biocontrol agents. Whole-genome analysis on seven selected strains confirmed the presence of numerous gene clusters with PGP and biocontrol functions and of novel secondary-metabolite biosynthetic genes, which could exert beneficial impacts on plant growth and protection. The predicted biocontrol potential was confirmed in dual culture assays against several phytopathogenic fungi and bacteria. Interestingly, the presence of predicted gene clusters with known biocontrol functions in some of the isolates was not predictive of the in vitro results, supporting the need of combining laboratory assays and genome mining in PGPB identification for future applications.