ArticlePDF Available

Abstract and Figures

Salt stress can affect survival, multiplication and ability of plant growth promoting microorganisms to enhance plant growth. Changes in a microbe’s proteome profile is one of the mechanisms employed by PGPM to enhance tolerance of salt stress. This study was focused on understanding changes in the exoproteome profile of Bacillus amyloliquefaciens EB2003A and Lactobacillus helveticus EL2006H when exposed to salt stress. The strains were cultured in 100 mL M13 ( B. amyloliquefaciens ) and 100 mL De man, Rogosa and Sharpe (MRS) ( L. helveticus ) media, supplemented with 200 and 0 mM NaCl (control), at pH 7.0. The strains were then incubated for 48 h (late exponential growth phase), at 120 rpm and 30 ( B. amyloliquefaciens ) and 37 ( L. helveticus ) °C. The microbial cultures were then centrifuged and filtered sterilized, to obtain cell free supernatants whose proteome profiles were studied using LC–MS/MS analysis and quantified using scaffold. Results of the study revealed that treatment with 200 mM NaCl negatively affected the quantity of identified proteins in comparison to the control, for both strains. There was upregulation and downregulation of some proteins, even up to 100%, which resulted in identification of proteins significantly unique between the control or 200 mM NaCl ( p ≤ 0.05), for both microbial species. Proteins unique to 200 mM NaCl were mostly those involved in cell wall metabolism, substrate transport, oxidative stress tolerance, gene expression and DNA replication and repair. Some of the identified unique proteins have also been reported to enhance plant growth. In conclusion, based on the results of the work described here, PGPM alter their exoproteome profile when exposed to salt stress, potentially upregulating proteins that enhance their tolerance to this stress.
Content may be subject to copyright.
Frontiers in Microbiology 01 frontiersin.org
Eect of NaCl stress on
exoproteome profiles of Bacillus
amyloliquefaciens EB2003A and
Lactobacillus helveticus EL2006H
JudithNaamala , SowmyalakshmiSubramanian ,
LeviniA.Msimbira and DonaldL.Smith *
Department of Plant Science, McGill University, Montreal, QC, Canada
Salt stress can aect survival, multiplication and ability of plant growth promoting
microorganisms to enhance plant growth. Changes in a microbe’s proteome
profile is one of the mechanisms employed by PGPM to enhance tolerance of
salt stress. This study was focused on understanding changes in the exoproteome
profile of Bacillus amyloliquefaciens EB2003A and Lactobacillus helveticus
EL2006H when exposed to salt stress. The strains were cultured in 100  mL  M13 (B.
amyloliquefaciens) and 100  mL De man, Rogosa and Sharpe (MRS) (L. helveticus)
media, supplemented with 200 and 0  mM NaCl (control), at pH 7.0. The strains
were then incubated for 48  h (late exponential growth phase), at 120  rpm and
30 (B. amyloliquefaciens) and 37 (L. helveticus) °C. The microbial cultures were
then centrifuged and filtered sterilized, to obtain cell free supernatants whose
proteome profiles were studied using LC–MS/MS analysis and quantified using
scaold. Results of the study revealed that treatment with 200  mM NaCl negatively
aected the quantity of identified proteins in comparison to the control, for both
strains. There was upregulation and downregulation of some proteins, even up to
100%, which resulted in identification of proteins significantly unique between the
control or 200  mM NaCl (p 0.05), for both microbial species. Proteins unique
to 200  mM NaCl were mostly those involved in cell wall metabolism, substrate
transport, oxidative stress tolerance, gene expression and DNA replication and
repair. Some of the identified unique proteins have also been reported to enhance
plant growth. In conclusion, based on the results of the work described here,
PGPM alter their exoproteome profile when exposed to salt stress, potentially
upregulating proteins that enhance their tolerance to this stress.
KEYWORDS
PGPM, Lactobacillus helveticus, Bacillus amyloliquefaciens, salt stress, exoproteome
profile
1. Introduction
Plant growth promoting microorganisms (PGPM) and their derivatives are key technology
sources for sustainable agriculture, especially with the urgent need to slow climate change and
its adverse eects (Naamala and Smith, 2020). For centuries, use of PGPM based inoculants to
sustainably enhance plant growth and increase yield, under stressed and ideal conditions has
been practiced in dierent parts of the world (Babalola and Glick, 2012; Bashan etal., 2014;
García-García etal., 2020; Naamala etal., 2023). e ability of PGPM and or their derivatives
OPEN ACCESS
EDITED BY
Woo-Suk Chang,
University of Texas at Arlington, UnitedStates
REVIEWED BY
Do-Won Jeong,
Dongduk Women’s University, Republic of
Korea
Pushp Sheel Shukla,
Dalhousie University, Canada
Klára Kosová,
Crop Research Institute, Czechia
*CORRESPONDENCE
Donald L. Smith
donald.smith@mcgill.ca
RECEIVED 17 April 2023
ACCEPTED 31 July 2023
PUBLISHED 28 August 2023
CITATION
Naamala J, Subramanian S, Msimbira LA and
Smith DL (2023) Eect of NaCl stress on
exoproteome profiles of Bacillus
amyloliquefaciens EB2003A and Lactobacillus
helveticus EL2006H.
Front. Microbiol. 14:1206152.
doi: 10.3389/fmicb.2023.1206152
COPYRIGHT
© 2023 Naamala, Subramanian, Msimbira and
Smith. This is an open-access article distributed
under the terms of the Creative Commons
Attribution License (CC BY). The use,
distribution or reproduction in other forums is
permitted, provided the original author(s) and
the copyright owner(s) are credited and that
the original publication in this journal is cited,
in accordance with accepted academic
practice. No use, distribution or reproduction is
permitted which does not comply with these
terms.
TYPE Original Research
PUBLISHED 28 August 2023
DOI 10.3389/fmicb.2023.1206152
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 02 frontiersin.org
to enhance plant growth is associated with their ability to exude in
their growth environment, proteins and metabolites with plant growth
promoting characteristics (Prithiviraj etal., 2003; Gray etal., 2006b;
Schwinghamer etal., 2016; Piechulla etal., 2017).
Salinity stress is a major global constraint to crop production,
aecting both plant yield quality and quantity. Although PGPM can
mitigate the eects of salinity stress on plants, it can also aect the
ability of PGPMs to enhance plant growth and may lead to microbial
death in case of exposure to levels beyond those tolerated (Zahran,
1997; Soussi etal., 2001; Nadeem etal., 2015; Naamala etal., 2022,
2023). Some microbes have developed mechanisms for surviving at
high salt concentrations. e ability of microbes to tolerate saline
conditions is in part dependent on their ability to regulate salt
concentration in their cytoplasm, in relation to that of their growth
environment. Mechanisms employed to regulate salt concentration
within the microbe include accumulation of osmolytes such as
glutamate and proline in their cytoplasm through de novo synthesis or
uptake from their growth environment (Zahran, 1997; Soussi etal.,
2001; Oren, 2008; Bojanovic etal., 2017), upregulation of iron uptake
mechanisms such as production of siderophores (Bojanovic etal.,
2017), alteration of their cell membrane composition (Bojanovic etal.,
2017; Hachicho et al., 2017), and maintenance of a high KCl
concentration in their cytoplasm to match that of their growth
medium (Oren, 2002; Oren, 2008). Eecting these mechanisms may
necessitate the microbe to make changes to its genome, proteome, and
metabolome proles, most probably upregulating those components
of each essential for enhancing salt tolerance mechanisms.
Protein expression occurs when genes are transcribed into
messenger RNA (mRNA), which is then translated into proteins, which
are major constituents of microbial cells (Karpievitch et al., 2010;
Zhang etal., 2010). Protein expression and secretion are usually in
response to either internal or external stimuli such as exposure to biotic
and abiotic stress (Zhang etal., 2010; Armengaud etal., 2012; Schoof
etal., 2022). e microbial proteome loosely translates to all proteins
associated with a given microbe. e microbial exoproteome refers to
proteins found in the immediate extracellular milieu of a microbe,
arising from active cellular secretion, passive excretion and or cell lysis
(Desvaux etal., 2010; Armengaud etal., 2012; Rubiano-Labrador etal.,
2015; Schoof et al., 2022). For microbes cultured in laboratories,
microbial exoproteome would refer to total proteins in spent media
aer removal of all microbial cells through centrifugation and ltration.
Exoproteome composition reects a microbes physiological state at a
given time and can provide insight into a microbes interactions with
its surroundings (Armengaud etal., 2012). Abiotic stresses such as
salinity, acidity and alkalinity aect the quantity and quality of proteins
synthesized and expressed by a microbe at a given time (Singleton
etal., 1982; Soussi etal., 2001; Msimbira etal., 2022). Exploring the
exoproteome of a microbe exposed to salt stress can provide insight
into a set of proteins expressed in response to salt stress, which could
then enhance our understanding of salt tolerance mechanisms in
microbes (Rubiano-Labrador etal., 2015). In general, the ‘omics
studies of biological systems have resulted in better understanding of
microbes and their environment (Karpievitch etal., 2010). Advances
in technology, such as invention of high through put tandem mass
spectrometry and liquid chromatography have allowed for easy
identication, analysis, classication, and function annotation of
complex protein samples (Listgarten and Emili, 2005; Zhang etal.,
2010; Armengaud, 2013; Kucharova and Wiker, 2014; Msimbira etal.,
2022). It is interesting to note that while some PGPM may lose their
ability to enhance plant growth following exposure to salt stress, others
may gain or bemore eective at enhancing plant growth, aer exposure
to some level of stress (Subramanian et al., 2021). erefore,
understanding how microbial exoproteome proles change with
changes in salt stress can improve utilization of CFS as plant growth
biostimulants, and enhance our elucidation of mechanisms employed
to enhance plant growth and or tolerate salt stress, given that some
proteins such as enzymes play a vital role in stress tolerance and plant
growth stimulation (Ahmad etal., 2010).
Bacillus amyloliquefaciens are rod shaped endospore forming gram
positive bacteria from the genus Bacillus and family Baciliaceae
(Woldemariam etal., 2020; Ngalimat etal., 2021). B. amyloliquefaciens is
widely used in the food, pharmaceutical and agricultural sectors
(Woldemariam etal., 2020). B. amyloliquefaciens and its derivatives have
been reported to enhance plant growth under stressed and ideal
conditions (Cappellari and Banchio, 2020; Cappellari et al., 2020; Duan
etal., 2021; Kazerooni etal., 2021; Naamala etal., 2022). Lactobacillus
helveticus is a gram positive facultative anaerobic lactic acid bacterium
(LAB) that is widely used in the food processing industry. However,
L. helveticus and its derivatives have also been reported to enhance plant
growth (Naamala etal., 2023). is study was focused on understanding
changes in exoproteome proles of B. amyloliquefaciens EB2003A and
L. helveticus EL2006H, at 0 mM NaCl and 200 mM NaCl. Results from
previous studies have shown that CFSs of both strains, when exposed to
200 mM NaCl, enhanced germination and radicle length of soybean, and
corn, as well as growth variables of potato (Naamala etal., 2022, 2023).
Wetherefore point out some of the proteins identied in this study, that
have been reported to enhance plant growth, as well as some of the
mechanisms plants employ to support growth.
2. Materials and methods
2.1. Obtaining protein samples
B. amyloliquefaciens EB2003A and L. helveticus EL2006H, which
were generously provided by EVL Inc., were cultured in 100 mL M13
and 100 mL De man, Rogosa and Sharpe (MRS) media, respectively,
supplemented with 200 and 0 mM NaCl (control), at pH 7.0. ey
were incubated for 48 h (late exponential growth phase), at 120 rpm
and a temperature of 30 and 37°C, for B. amyloliquefaciens EB2003A
and L. helveticus EL2006H, respectively. Four replicates per treatment
per species were cultured. Microbial cultures were then centrifuged,
using a Sorvall Biofuge Pico (Mandel Scientic, Guelph, ON, Canada),
for 10 min, at 10,000 rpm (15,180× g; SLA-1500) and 4°C, to pellet the
microbial cells and separate them from the cell-free supernatant (CFS)
(Gray etal., 2006a; Subramanian etal., 2021). e CFS was further
vacuum ltered using 0.22 μm nylon lters to ensure that all bacterial
cells were removed. Trichloroacetic acid (TCA; T9151, Sigma Aldrich)
precipitation was used to extract total proteins from the obtained CFS
replicates. e CFS was mixed with 100% (w/v) TCA, in 250 mL
conical asks to create a 25% working solution of TCA. e mixture
was incubated at 20°C for 1 h, and transferred to an orbital shaker
(MBI, Montreal Biotech Inc., Canada) with shaking speed of 90 rpm,
in a cold room, at a temperature of- 4°C, for protein precipitation,
overnight. is was followed by a 10 min centrifugation at 4°C and
10,000 rpm, to pellet the protein. e obtained protein pellet was then
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 03 frontiersin.org
washed with ice-cold acetone, air-dried under a laminar ow hood,
and dissolved in 2 M urea (U4883, Sigma Aldrich). e protein
obtained from the four replicates of each treatment were pooled to
form one sample per species. e experiment was repeated four times
to get the appropriate biological replicates. Concentration of proteins
obtained from the four experiments was determined, using the Lowry
method (Lowry etal., 1951).
2.2. LC–MS/MS protein profiling
Aer determining the concentration of the obtained proteins,
10 μg protein per sample was dissolved in 20 μL of 2 M urea and sent
to Montreal Clinical Research Institute (IRCM), for liquid
chromatography mass spectrometry (LC–MS/MS) analysis. Total
proteins were digested using trypsin enzyme and injected into an LC–
MS/MS equipped with Linear Trap Quadrupole Velos Orbitrap
(ermo Fisher Waltham, MA, UnitedStates). e data set obtained
from the mass spectra were searched against Bacillus spp. and
Lactobacillus spp. databases, using Mascot soware (Matrix Science,
London, UnitedKingdom). Scaold Soware (version 5.1.2, Proteome
Soware Inc., Portland OR) was used to validate the obtained MS/MS
based peptides and proteins, using an equal to or greater than 95%
acceptance of protein probability, with a minimum of two peptides
and 95% peptide probability (Keller etal., 2002).
2.3. Quantitative data analysis
Proteomic data for identied proteins obtained from the LC–MS/
MS analysis was quantitatively analyzed, based on spectra count values,
using Scaold 5 (Scaold Soware for MS/MS Proteomics). Spectra
count values were normalized and subjected to analysis of variance, at
the 5% signicance level, using a Benjamini-Hochberg multiple test
correction, to detect signicant dierences between treatments.
Signicance was based on both sher’s exact test (p 0.05) and fold
change of more than or equal to 1.2. FASTA les generated from
Scaold 5 were analyzed using OmicsBox for functional annotation
and interpretation of the protein sequences. Volcano plots were created
using OriginPro soware (OriginPro learning edition, version 2023
learning edition) while Venn diagrams were generated using Scaold
soware for MS/MS proteomics. e LC–MS/MS proteomic data are
available in the Mass Spectrometry Interactive Virtual Environment
(MassIVE) at doi:10.25345/C5PG1HZ4M and PXD041778 for Bacillus
amyloliquefaciens EB2003A, and doi:10.25345/C54B2XF6V and
PXD041177 for Lactobacillus helveticusel2006h.
3. Results
3.1. Exoproteome analysis for Bacillus
amyloliquefaciens EB2003A
Based on scaold and OmicsBox analyzes of the LC–MS data,
there were variations in identied proteins for CFS of
B. amyloliquefaciens EB2003A cultured at 0 mM NaCl and 200 mM
NaCl, as shown in Table1 and Supplementary Table S1. In general,
NaCl lowered the quantity of identied proteins, total unique spectra,
and total unique peptides, as shown in Figure1, visualized at 95%
protein threshold, 2 minimum peptides and 0.00% decoy FDR. A total
of 1,295 proteins, 9,718 total peptides and 15,283 total spectra were
identied. Out of the observed proteins, 1,024 were shared between
both salt levels while 197 were unique to 0 mM NaCl, and 74 proteins
were unique to 200 mM NaCl.
Further quantitative analysis of the LC–MS/MS data output using
scaold showed a signicant decrease in the quantity of identied
proteins at 200 mM NaCl in comparison to 0 mM NaCl, at p 0.05
(Fisher’s exact test). Several proteins were upregulated or down
regulated at both salt levels as shown in Supplementary Table S1.
Likewise, a number of proteins were unique to either 0 or 200 mM
NaCl. Analysis with a 1.2 fold change also showed signicant
variations in proteins identied for 0 and 200 mM NaCl, as shown in
Figure 2. Supplementary Table S3 shows OmicsBox data for
B. amyloliquefaciens EB2003A.
3.2. Exoproteome analysis for Lactobacillus
helveticus EL2006H
Based on scaold and OmicsBox analyzes of the LC–MS/MS data,
there were variations in identied proteins for L. helveticus EL2006H
cultured at 0 mM NaCl and 200 mM NaCl, as shown in Table2 and
Supplementary Table S2. Two hundred mM NaCl greatly aected
identied proteins, with the majority downregulated, even to 100%.
Figure3 shows a comparison of the quantity of total proteins, total
unique peptides, and total unique spectra for 0 and 200 mM NaCl,
visualized at 95% protein threshold, 2 minimum peptides and 0.00%
decoy FDR. A total of 317 proteins, 1,628 peptides and 2,307 spectra
were observed. Out of the observed proteins, 136 were shared between
both salt levels while 178 were unique to 0 mM NaCl, and 3 were
unique to 200 mM NaCl, as shown in Figure3.
Further quantitative analysis of the LC–MS/MS output, using
scaold showed a signicant decrease in identied proteins at 200 mM
NaCl, in comparison to 0 mM NaCl, at p 0.05 (Fishers exact test).
e majority of the proteins were signicantly downregulated at
200 mM NaCl (Supplementary Table S2), with only seven upregulated
proteins, namely, a cluster of hypothetical proteins GFB61_00500, a
cluster of SLAP domain-containing proteins, a cluster of peptide ABC
transporter substrate-binding proteins, surface proteins, bronectin
type III domain-containing proteins, a cluster of Stk1 family PASTA
domain containing Ser/r kinases, and a cluster of metal ABC
transporter substrate binding proteins. Based on the fold analysis (1.2-
fold change and above), only three proteins namely, bronectin type
III domain-containing protein, cluster of hypothetical proteins and
cluster of metal ABC transporter, were signicantly upregulated as
shown in Supplementary Table S2. Supplementary Table S4 shows
OmicsBox data for L. helveticus 2006H. Figure4 is a volcano plot
illustrating the distribution of identied proteins for
L. helveticus EL2006H.
3.3. Functional annotation of proteins of
Bacillus amyloliquefaciens CFS observed at
0 and 200  mM NaCl
Using gene ontology (GO) enrichment analysis, identied
proteins were grouped into four groups, based on their functions,
namely, enzyme code distribution, cellular components, biological
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 04 frontiersin.org
processes, and molecular functions, as shown in Table1. ere was a
variation in the eect of 200 mM NaCl on the proteins performing
various functions, majority being downregulated while some were
upregulated, yet others were unique to 200 mM NaCl. Worthy noting
is that all proteins involved in the various listed biological processes
were downregulated while proteins related to the extracellular region
cellular component were unique to 200 mM NaCl. Proteins involved
in catalyzing proteins and nucleic acids were also upregulated by 18.8
and 17.5%, respectively. Under the enzyme code distribution, all
enzymes were downregulated except for Translocases which were
upregulated by 20.6%.
3.4. Functional annotation of proteins of
Lactobacillus helveticus EL2006H CFS
observed at 0 and 200  mM NaCl
Using gene ontology (GO) enrichment analysis, identied
proteins were grouped in four sets, namely, biological processes,
cellular components, molecular functions, and enzyme code
distribution, as shown in Table2. ere was a variation in the eect of
200 mM NaCl on the proteins performing various functions; the
majority were downregulated although a few were upregulated.
Proteins in some functional groups were unique to 200 mM NaCl
(Table2). For example, under biological processes, transmembrane
transport and cell adhesion were unique to 200 mM NaCl while under
cellular components, transporter complex and membrane protein
complex were unique to 200 mM NaCl. Under molecular function,
peptidoglycan muralytic activity, protein binding and structural
constituents of cell walls were unique to 200 mM NaCl. Notably, all the
upregulated/unique proteins perform functions related to cell wall
metabolism or substrate transportation.
4. Discussion
Uncontrollable changes in microbial environments, especially
under eld conditions, requires PGPM to adapt to the changes in
order to survive (Gao etal., 2007; Galperin, 2010). Salinity stress is a
leading global abiotic stress aecting crops and PGPM proliferation
(Liu etal., 2023). When a microbe is exposed to stress, it may alter its
proteome prole, upregulating proteins essential for enhancing
tolerance to the stress while down regulating those that are likely not
so essential (Galperin, 2010; Msimbira et al., 2022). As a result,
proteome proles of a microbe grown in dierent environmental
conditions may vary signicantly. e current study compared
exoproteome proles of B. amyloliquefaciens EB2003A and
L. helveticus EL2006H exposed to 0 and 200 mM NaCl. Results of the
study showed variations in total proteins identied for both strains at
the two salt levels. Some of the identied proteins were unique to
either 0 or the 200 mM NaCl. Among the proteins unique to 200 mM
NaCl were cell wall metabolic enzymes, transcription/translation
regulators, potential virulence factors, phage proteins, antibiotics
resistance proteins, solute transporter proteins and, of course,
hypothetical proteins. ese ndings are to some extent like those of
Pumirat et al. (2009) and Rubiano-Labrador et al. (2015) who
examined the exoproteome of Burkholderia pseudomallei and Tistlia
consotensis exposed to salinity stress.
e cell-wall is the outermost layer of a bacterial cell, that acts as
a stress barrier and maintains cell shape (Mueller and Levin, 2020).
erefore, maintaining the integrity of the cell wall is a mechanism for
stress tolerance in bacteria. is may explain why, based on GO
function analysis, the dierent functional groups identied in the
current study, protein classes performing functions related to the cell
wall and the extracellular region were upregulated at 200 mM NaCl,
in both B. amyloliquefaciens EB2003A and L. helveticus
EL2006H. Peptidoglycan is the major component of the cell wall,
TABLE1 Comparing the distribution of B. amyloliquefaciens EB2003A
proteins to the dierent functional groups according to GO enrichment
analysis, at 0  mM NaCl (control) and 200 mM NaCl (treatment).
# Sequences
Functional group 0  mM NaCl 200  mM NaCl
Biological process
Organic substance metabolic process 1,202 1,028 (14.5%)
Primary metabolic process 1,061 919 (13.4%)
Cellular metabolic process 1,100 912 (17.1%)
Nitrogen compound metabolic process 939 852 (9.3%)
Biosynthetic process 637 493 (22.6%)
Small molecule metabolic process 631 490 (22.3%)
Catabolic process 204 193 (5.4%)
Cellular components
Intracellular anatomical structure 500 395 (21%)
Cytoplasm 468 363 (22.4%)
Membrane 162 232 (30.2%)
Cell periphery 99 164 (39.6%)
Intrinsic component of membrane 110 159 (30.9%)
Extracellular region 0 19 (100%)
Molecular function
Ion binding 821 706 (14.0%)
Organic cyclic compound binding 701 620 (11.6%)
Heterocyclic compound binding 701 620 (11.6%)
Hydrolase activity 663 579 (12.7%)
Small molecule binding 580 463 (20.2%)
Oxidoreductase activity 438 331 (24.4%)
Transferase activity 481 354 (26.4%)
Carbohydrate derivative binding 350 296 (15.4%)
Catalytic activity, acting on a protein 177 218 (18.8%)
Catalytic activity, acting on a nucleic acid 160 194 (17.5%)
Ligase activity 54 0 (100%)
Enzyme code distribution
Hydrolases 618 537 (13.1%)
Isomerases 114 90 (21.1%)
Ligases 182 148 (18.7%)
Lyases 134 102 (23.9%)
Transferases 463 344 (25.7%)
Translocases 54 68 (20.6%)
Oxidoreductases 418 310 (25.8%)
e bold values represent proteins upregulated at 200 mM NaCl.
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 05 frontiersin.org
whose synthesis, polymerization, modication and turn over
contribute to maintaining cell wall integrity (Popham and Young,
2003; Sauvage etal., 2008; Shin etal., 2020). In this study, proteins,
such as LytR family transcriptional regulator, Amidases, peptidoglycan
endopeptidases and penicillin binding proteins (PBPs), SLAP domain-
containing protein and surface proteins were uniquely produced by
either B. amyloliquefaciens EB2003A or L. helveticus EL2006H,
exposed to 200 mM NaCl. ey all play vital roles in peptidoglycan
metabolism and maintenance of the cell wall. For example, LytR
family transcriptional regulator proteins, also known as LytR-
CpsA-Psr (LCP) family proteins, are in fact enzymes involved in the
attachment of Glycopolymers, such as wall teichoic acids on the
peptidoglycan (Kawai etal., 2011; Molloy, 2011; Gale etal., 2017;
Siegel etal., 2019). Previously, this family of proteins was reported to
play a transcription regulation role (Gao etal., 2007; Galperin, 2010),
although in later studies, Kawai and co-authors disagreed (Kawai
etal., 2011), suggesting that these regulatory roles could bebacterial
genus, species, or strain specic. Bacterial amidases play a vital role in
bacteria cell wall metabolism because, especially under stressful
conditions, they are involved in the remodeling, turnover, recycling,
and metabolism of peptidoglycan (Park, 1995; Weber etal., 2013;
Senzani etal., 2017; Mueller and Levin, 2020). Endopeptidases play a
major role in maintaining bacterial cell integrity and shape, through
processes such as peptidoglycan turnover and modication (Shin
et al., 2020). Penicillin binding protein PBP4, a peptidoglycan
endopeptidase was reported to enhance tolerance of B. subtilis to salt
stress by modifying the peptidoglycan (Palomino et al., 2009).
B. subtilis was reported to recycle its peptidoglycan toward the end of
its exponential growth, entering stationary phase, which could enable
prolonged survival of the bacteria during stationary phase (Borisova
etal., 2016). Surface proteins, also known as the glycoprotein layer or
S layer proteins (Engelhardt, 2007; Hynönen and Palva, 2013), and
surface layer associated proteins (SLAP), in the current study, were
unique to L. helveticus treated with 200 mM NaCl. Expression of
surface proteins has been linked to the ability of some Lactobacilli
species to tolerate changes in the human gastrointestinal tract
conditions, such as bile and low pH (Sengupta etal., 2013). It should
benoted that, not all prokaryotes produce surface proteins and that
their role varies from one group to another, leaving no universal
function of surface proteins in species that do possess them
(Engelhardt, 2007). Upregulation of surface proteins was also observed
in Lactobacillus acidophilus IBB 801 exposed to dierent abiotic
stresses such as NaCl, bile salt and high temperature (Grosu-Tudor
etal., 2016). Deletion of IgdA SLAP resulted in a mutant that was more
sensitive to salt stress and had a visibly disrupted cell surface when
compared to strains with the protein (Klotz etal., 2020). However, the
mechanisms through which surface proteins and SLAP enhance
tolerance to salt stress is yet to beveried. Likewise, Stk1 family
PASTA domain-containing Ser/r kinase also upregulated in
L. helveticus exposed to 200 mM NaCl have been reported to play a
role in bacteria cell wall metabolism and cell division (Janczarek etal.,
2018). Its expression was reported to enhance tolerance of
Streptococcus suis serotype 2 to oxidative stress (Zhu etal., 2014).
Exposure to stress may bea trigger for bacteria to reprogram their
gene expression, consequently resulting in new gene products that
could beessential for stress tolerance. In the current study, proteins
such as MarR family transcriptional regulators and rRNA
pseudouridine synthase were upregulated at 200 mM NaCl. e MarR
FIGURE1
Comparison of total proteins, total peptides and total spectra identified at 0 and 200  mM NaCl, for B. amyloliquefaciens EB2003A (p 0.05).
FIGURE2
Volcano plots showing the distribution of identified proteins for B.
amyloliquefaciens EB2003A as log10 (Benjamini–Hochberg-adjusted
p-values) plotted against log2 (fold change) for 0 vs. 200 mM NaCl.
The two blue dotted vertical lines represent a ± 1.2-fold change, while
the pink dotted horizontal line indicates the significance threshold
(before logarithmic transformation) p 0.02425.
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 06 frontiersin.org
family transcriptional regulators constitute a prominent family of
transcription factors involved in the reprogramming of gene
expression in response to stress conditions, such as oxidative stress
(Pérez-Rueda etal., 2004; Grove, 2013; Deochand and Grove, 2017).
MarR are involved in metabolism and antibiotic resistance of some
bacteria (Will and Fang, 2020). e enzyme rRNA pseudouridine
synthase catalyzes the synthesis of RNA pseudouridine from uracil,
the most common modied nucleoside in rRNA that plays a role in
gene expression (Ofengand, 2002; Zhao et al., 2018). Sulfate
Transporter and Anti-Sigma factor antagonist (STAS) domain-
containing proteins are produced in multiple species, including
bacteria and mammals (Moy and Seshu, 2021). In bacteria, they are
associated with stress tolerance among other factors, by regulating the
large family of sigma factors (ρ) that bind to RNA polymerase to
confer transcriptional target gene specicity (Moy and Seshu, 2021).
For instance, sporulation in B. subtilis, which is a response to stress
involves anti-anti-sigma factors, or anti-anti-ρ, which are STAS
domain proteins (Sharma etal., 2011). Changes in gene products may
also aect cellular biochemical composition and cellular processes, as
observed in the current study.
Salt stress leads to osmotic, oxidative, and ionic stress in microbes,
which, depending on their severity could cause damage to cellular
components such as the cell membrane, nucleic acids, enzymes, and
other proteins. A high concentration of salt ions such as Na+ and Cl
in the microbe’s extracellular environment may cause loss of water
from the microbial cell, leading to loss of cell turgor pressure as well
as reduction in cellular processes such as metabolism (Krämer, 2010;
Tsuzuki etal., 2011; Rubiano-Labrador etal., 2015). To counter act
such eects, microbes develop mechanisms that enhance microbial
tolerance to oxidative, osmotic, or ionic stress. In the current study, at
200 mM NaCl, proteins that enhance bacterias tolerance to oxidative
stress were unique to 200 mM NaCl. For example, ioredoxins play
a major role in bacterial response to oxidative stress by quenching
singlet oxygen, scavenging hydroxyl radicals and donating hydrogen
to peroxidases (Chae etal., 1994; Das and Das, 2000; Zeller and Klug,
2006; Lu and Holmgren, 2014; Cheng etal., 2017). Members of the
xenobiotic response element (XRE) family transcriptional regulators,
among other functions, have been reported to enhance oxidative stress
tolerance in dierent bacteria species such as Streptococcus suis and
Corynebacterium glutamicum (Hu etal., 2019; Si etal., 2020; Zhang
etal., 2022). Flavodoxin family proteins were reported to enhance
tolerance of plant growth promoting rhizobacteria, such as
Pseudomonas uorescens Aur6 and Ensiifer meliloti, to oxidative stress
(Coba de la Peña etal., 2013). Heme A synthase catalyzes the synthesis
of heme A from heme O (Lewin and Hederstedt, 2016). Heme A is
particularly a co-factor of terminal oxidase enzymes involved in
oxygen reduction during aerobic respiration (Hederstedt, 2012; Choby
# Sequences
Functional group 0  mM NaCl 200  mM NaCl
Isomerases 33 3 (90.1%)
Translocases 9 6 (33.3%)
Lyases 18.5 0 ( 100%)
Ligases 21.75 0 (100%)
e bold values represent proteins upregulated at 200 mM NaCl.
TABLE2 (Continued)TABLE2 Comparing the distribution of L. helveticus EL2006H CFS
proteins to the dierent functional groups according to GO enrichment
analysis, at 0  mM NaCl (control) and 200 mM NaCl (treatment).
# Sequences
Functional group 0  mM NaCl 200  mM NaCl
Biological process
Organic substance metabolic process 320 50 (84.4%)
Primary metabolic process 294 44 (85.0%)
Nitrogen compound metabolic process 282 46 (83.6%)
Cellular metabolic process 239 12 (94.9%)
Biosynthetic process 149 0 (100%)
Small molecule metabolic process 76 5 (93.4%)
Catabolic process 60 4 (93.3%)
Establishment of localization 62 51 (17.7%)
ATP metabolic process 8 0 (100%)
Transmembrane transport 0 40 (100%)
Cell adhesion 0 7 (100%)
Cellular components
membrane 170 94 (44.7%)
intracellular anatomical structure 161 0 (100%)
intrinsic component of membrane 147 81 (44.8%)
cell periphery 110 75 (31.8%)
cytoplasm 92 0 (100%)
organelle 72 0 (100%)
extracellular region 17 24 (29.2%)
external encapsulating structure 29 24 (17.2%)
Transporter complex 0 19 (100%)
Membrane protein complex 0 19 (100%)
Molecular function
Hydrolase activity 189 76 (59.7%)
Organic cyclic compound binding 200 7 (96.5%)
Heterocyclic compound binding 200 7 (96.5%)
Ion binding 138 14 (89.8%)
Catalytic activity, acting on a protein 89 39 (56.2%)
Small molecule binding 89 3 (96.6%)
Transferase activity 86 0 (100%)
Structural constituent of ribosome 67 0 (100%)
Carbohydrate derivative binding 53 0 ( 100%)
Isomerase activity 9 0 (100%)
ATP hydrolysis activity 18 0 (100%)
Peptidoglycan muralytic activity 0 9 ( 100%)
Protein binding 0 16 ( 100%)
Structural constituent of cell wall 0 6 ( 100%)
Enzyme code distribution
Oxidoreductases 38 4 (89.5%)
Transferases 86 4 (95.3%)
Hydrolases 167 76 (54.5%)
(Continued)
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 07 frontiersin.org
and Skaar, 2016). High levels of intracellular heme have been reported
to activate Hap1p which subsequently induces the transcription of
genes involved in oxidative stress response (Martínez etal., 2016).
Since salt stress can result in damage to essential microbial
constituents such as enzymes, and nucleic acids, DNA and RNA, it is
important that damaged components are repaired or replaced with
new ones. Upregulation of proteins that are directly or indirectly
involved in synthesis and repair of cellular proteins and nucleic acids
was observed in the current study. e enzyme 2,3-cyclic-nucleotide
2-phosphodiesterase plays a major role in the metabolism of purines
and pyrimidines, the building blocks of DNA and RNA, and provide
energy and co-factors important in cell-division (Yin etal., 2018). is
is because it contains cyclic phosphodiesterase and 3-nucleotidase
activity and catalyzes the hydrolysis of 2,3-cyclic nucleotides to yield
nucleotides and phosphate. erefore, the enzyme plays a role in DNA
and RNA synthesis and repair through provision of building blocks.
e enzyme m
1
A22-tRNA methyltransferase (TrmK) catalyzes N
(1)-adenosine methylation to N1 of adenine 22 of bacterial tRNA
(Roovers etal., 2008; Sweeney etal., 2022). Addition of a methyl group
plays a role in maintaining stability of tRNA (Roovers etal., 2008).
Stability of tRNA is essential in protein synthesis since they bridge the
gap between mRNA and amino acids during translation. e enzyme
thioredoxin plays a major role in protein repair and DNA synthesis by
donating hydrogen that reduces ribonucleotide reductase and
methionine sulfoxide reductase which catalyze the process (Zeller and
Klug, 2006; Lu and Holmgren, 2014). e upregulation of such
enzymes may also explain the high frequency of protein classes whose
function annotation involves catalytic activity, acting on nucleic acid,
and catalytic activity acting on proteins, that was observed in
B. amyloliquefaciens EB2003A, at 200 mM NaCl.
When exposed to stress, its paramount that microbes maintain an
even ow of substrates from their environment to the inside of the cell,
and vice versa. is ensures availability of carbon sources for energy as
well as metabolites required to serve purposes such as osmoregulation,
enzyme co-factors, synthesis of proteins and nucleic acids. e microbe
requires energy for channeling to survival mechanisms (Yan etal.,
2015; Msimbira et al., 2022). Microbes respond to osmotic stress
through intracellular accumulation of inorganic ions such as K+ and
organic solutes such as proline (Soussi etal., 2001; Bojanovic etal.,
2017). Although some of these osmo-protectants can besynthesized de
novo, it is less energy ecient than sourcing them from outside of the
cell (Zahran, 1997; Oren, 2008; Zhang etal., 2015; Bojanovic etal.,
2017; Lycklama etal., 2018). erefore, maintaining adequate substrate
transport systems is essential for microbial tolerance to stress. In this
study, weobserved upregulation of a number of proteins associated
with various substrate transport systems, such as the ATP binding
cassette (ABC) transporter, the major facilitator superfamily (MFS)
transporter and the phosphotransferase system (PTS) fructose
transporter subunit IIC in B. amyloliquefaciens EB2003A and
L. helveticus EL2006H. e ABC transporter facilitates transportation
of a wide range of substrates such as sugars, amino acids, and metals
from the microbe’s external environment (Du etal., 2011; Lycklama
et al., 2018; Teichmann et al., 2018). Among the observed ABC
transporter proteins were the amino acid ABC transporter substrate-
binding protein, multispecies: ABC transporter substrate-binding
protein, and multispecies: zinc ABC transporter substrate-binding
protein. e MFS transporter is one of the oldest protein families, a
group of secondary active transporters involved in selective
FIGURE3
Comparison of total proteins, total peptides and total spectra identified at 0 and 200  mM NaCl for L. helveticus EL2006H (p 0.05).
FIGURE4
Volcano plots showing the distribution of identified proteins for L.
helveticus EL2006H as log10 (Benjamini–Hochberg-adjusted
p-values) plotted against log2 (fold change) for 0 vs. 200  mM NaCl.
The two dotted black vertical lines represent a  ±  1.2-fold change,
while the blue dotted horizontal line indicates the significance
threshold (before logarithmic transformation) p 0.01935.
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 08 frontiersin.org
transportation of substrates such as carbohydrates, amino acids, and
lipids, to ions, peptides and nucleosides, across microbial membranes
and plays a role in other microbial physiological processes, such as
resistance to toxic compounds like antibiotics and salicylic acid, and
enhanced salt tolerance (Yan, 2013; Lee etal., 2016; Pasqua etal., 2019).
For instance, MFS eux pumps VceCAB were reported to enhance the
tolerance of E. coli to bile salts (Woolley etal., 2005). e PTS system
is involved in uptake and phosphorylation of carbohydrates as well as
signal transduction (Bernhard, 2012). e enzyme IIC component
selectively transports sugar molecules across microbial membranes
(Jahreis etal., 2008; Jason etal., 2014). is allows microbes such as
bacteria to eciently utilize carbohydrate sources of their choice, at a
given time (Jahreis etal., 2008). e fructose family is a subfamily and
the oldest of the glucose superfamily of PTS. e ability of an organism
to utilize various carbon sources enables them to survive in varying
environmental conditions. As a result, upregulation of sugar uptake
systems has been linked to microbe response to stress, because
microbes require nutrients and osmo-protectants for survival under
stressful conditions (Pittman etal., 2014).
When exposed to stress, some members of the genus Bacillus,
B. amyloliquefaciens EB2003A inclusive, form spores which are
essential for survival under stressful conditions for long periods of
time (Setlow, 2014; Ghosh etal., 2018). Once favorable conditions are
restored, the spores germinate, giving rise to new microbial cells. e
germination of spores is triggered by amino acids such as L-alanine,
L- valine and L- asparagine (Setlow, 2014; Ghosh etal., 2018). In the
current study, alanine containing proteins: cation symporter family
protein and asparagine synthetase B were unique to the 200 mM NaCl
exoproteome of B. amyloliquefaciens EB2003A. e enzyme
asparagine synthase catalyzes the synthesis of asparagine from
aspartate and glutamine (Lomelino etal., 2017; Zhu etal., 2019). e
alanine cation symporter family protein is a transporter protein that
transports alanine but no other amino acids (Ma etal., 2019).
In addition to the proteins with known functions, several
hypothetical proteins were also unique to 200 mM NaCl treatment.
Proteins are classied as hypothetical if a corresponding mRNA
sequence is available in the data base, but there is no similar protein
sequence, hence, insucient information concerning their possible
functions. However, it’s possible that such proteins play a role in
enabling the microbes survival in growth conditions under which
they are produced.
In previous studies, CFS of B. amyloliquefaciens EB2003A and
L. helveticus EL2006H exposed to 200 mM NaCl enhanced
germination and radicle length of corn and soybean and growth
variables of potato grown under NaCl stress and normal conditions
(Naamala etal., 2022, 2023). e ability of a microbe to enhance
plant growth is related to its ability to exude into their growth
environment, bioactive substances with ability to enhance plant
growth. Among the proteins upregulated at 200 mM NaCl, in the
current study, are those that have been reported to enhance plant
growth under stressed and ideal conditions. Whether these proteins
were in part responsible for the bioactivity observed in our previous
study needs to beinvestigated further. However, application of
exogenous heme has been reported to enhance plant tolerance to
stress such salt stress (Woodson etal., 2011; Zhang etal., 2013; Wu
etal., 2022). e heme precursor 5-aminolevulinic acid (ALA) was
reported to enhance growth of plants exposed to salt stress (Hui
etal., 2006; Daneshmand and Oloumi, 2015; Genişel and Erdal,
2016; Wu etal., 2022). Exogenous application of ALA, resulted in
an increase in heme content, an indication that heme is involved in
the role of ALA in alleviating salinity stress (Wu etal., 2022). Heme
is involved in the transformation of superoxide anions in the
antioxidant system, hence, potentially playing a pivotal role in
mitigating the eects of oxidative stress on plants (Wu etal., 2022).
Esterases have been reported to play a role in plant growth and
development, involved in such crucial stages as seed germination,
pollen development, lateral root, and overall root development
(Takahashi etal., 2010; Clauss etal., 2011; Dolui and Vijayaraj,
2020; Zhang etal., 2020; Ursache etal., 2021; Shen etal., 2022). In
fact, esterases are believed to have played a role in the evolutionary
colonization of land by plants, through the conservation of water in
a desiccating environment (Niklas etal., 2017; Philippe etal., 2020).
MarR homologs were involved in symbiotic plant microbe
interactions. For example, the MarR homolog ExpG Sinorhizobium
meliloti activates transcription of three exp. operons that are
involved in the production of galactoglucan, which it needs for
plant root nodulation (Becker etal., 1997; Bartels et al., 2003).
Proteins such as the MFS eux pumps were reported to beinvolved
in the interaction of plants and symbiotic microbes, such as rhizobia
through enhancing nodulation and enhancing tolerance to
avonoids (Pasqua et al., 2019). ioredoxin, another protein
(enzyme) that was unique to 200 mM NaCl exoproteome of
B. amyloliquefaciens EB2003A is also found in higher plants where
it is classied as a disulde regulatory protein, belonging to a
complex of regulatory proteins consisting of types f, m, x, y, h, and
o (Meng etal., 2010). ioredoxin proteins play major roles in the
regulation of carbon metabolism, embryogenesis, chloroplast
development and mobilization of seed reserves, in plants (Jedelská
etal., 2020). ey also play a role in plant responses to biotic and
abiotic stresses through protection from reactive oxygen species
(Dos Santos and Rey, 2006; Meyer etal., 2012; Geigenberger etal.,
2017). ioredoxin h ortholog Trx h9, was reported play a role in
the germination of wheat (Li etal., 2009). It is believed that Trx h
regulates seed germination by reducing the disulde proteins stored
in the dry seed to the sulydryl state, following the addition of
water (Maeda etal., 2003, 2005; Rhazi etal., 2003; Zahid et al.,
2008). Meng et al. (2010) observed chlorotic leaves, short and
smaller roots in Arabidopsis thaliana Trx h9 mutants. Furthermore,
mutant plants were dwarf, with small and irregular mesophyll cells,
as well as lower chloroplast numbers and less chlorophyll, in
comparison to the wild type control, suggesting that Trx h plays a
role in plant growth (Meng et al., 2010). High expression of
thioredoxin h8 was observed in tobacco plants whose growth was
enhanced when treated with Bacillus aryabhattai (Xu etal., 2022).
Amidases are involved in the biosynthesis of indole acetic acid, a
phytohormone that plays major roles in plant growth and
development (Spaepen et al., 2007). Other PGPM such as
Pseudomonas putida have also been reported to produce amidase
(Chacko et al., 2009). Because they are involved in nitrogen
metabolism, amidases increase nitrogen use eciency in plants,
which subsequently enhances plant growth under both stressed and
non-stressed conditions (Unkefer etal., 2023).
ere are several mechanisms through which the identied
proteins can enhance plant growth. ese include regulation of the
anti-oxidant system, regulation of the photosynthetic system, ion
balance, hydrolysis of compounds that aect plant quality,
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 09 frontiersin.org
mobilization of nutrients during seed germination, biosynthesis of
phytohormones involved in metabolic pathways such as nitrogen
metabolism and maintenance of plant fertility (Spaepen etal., 2007;
Takahashi etal., 2010; Clauss etal., 2011; Dolui and Vijayaraj, 2020;
Zhang etal., 2020; Unkefer etal., 2023).
Based on these studies, it’s possible that some of the proteins
upregulated at 200 mM NaCl stress were responsible for enhancing
radicle length, germination of corn and soybean, and growth variables
of potato, as observed in our previous studies.
5. Conclusion
Salinity stress aects survival, growth, and ability of PGPM to
enhance plant growth. However, some PGPM have developed
mechanisms of tolerating high levels of salt, altering their exoproteome
prole being one. In the current study, B. amyloliquefaciens EB2003A
and L. helveticus EL2006H, exhibited unique proteins when exposed
to 200 mM NaCl, some of which have also been reported to enhance
plant growth. Results of the study are in line with previous reports that
when exposed to stress, microbes alter their exoproteome prole. To
the best of our knowledge, this is the rst study to report on the eect
of NaCl on B. amyloliquefaciens EB2003A and L. helveticus EL2006H
exoproteome proles. Findings of this study will expand knowledge
regarding mechanisms through which Bacillus spp. and Lactobacillus
spp. tolerate salt stress at the protein level.
Data availability statement
e datasets presented in this study can be found in online
repositories. e names of the repository/repositories and accession
number(s) can be found in the Mass Spectrometry Interactive Virtual
Environment (MassIVE) online repository, at doi:10.25345/
C5PG1HZ4M and PXD041778 for Bacillus amyloliquefaciens
EB2003A, and doi:10.25345/C54B2XF6V and PXD041177 for
Lactobacillus helveticusEL2006H.
Author contributions
JN set up the experiments, was involved in data analysis and wrote
the manuscript. SS advised on experimental set up, data analysis, and
manuscript editing. LM helped with data analysis and manuscript
review and editing. DS provided funding, the intellectual environment
and did extensive manuscript editing. All authors have read and
agreed to the nal version of the manuscript.
Funding
is work was funded through a grant from Consortium de
recherche et innovations en bioprocédés industriels au Québec,
number CRIBIQ 2017-034-C30, with support from synagri and
EVL inc.
Conflict of interest
e authors declare that the research was conducted in the
absence of any commercial or nancial relationships that could
beconstrued as a potential conict of interest.
Publisher’s note
All claims expressed in this article are solely those of the authors and
do not necessarily represent those of their aliated organizations, or
those of the publisher, the editors and the reviewers. Any product that
may be evaluated in this article, or claim that may be made by its
manufacturer, is not guaranteed or endorsed by the publisher.
Supplementary material
e Supplementary material for this article can befound online
at: https://www.frontiersin.org/articles/10.3389/fmicb.2023.1206152/
full#supplementary-material
SUPPLEMENTARY TABLE S1
Comparison of exoproteome for B. amyloliquefaciens EB2003A, at 0 and
200 mM NaCl.
SUPPLEMENTARY TABLE S2
Comparison of exoproteome for L. helveticus El2006H, at 0 and 200
mM NaCl.
SUPPLEMENTARY TABLE S3
B. amyloliquefaciens EB2003A OmicsBox data.
SUPPLEMENTARY TABLE S4
L. helveticus EL2006H OmicsBox data.
References
Ahmad, P., Jaleel, C. A., Salem, M. A., Nabi, G., and Sharma, S. (2010). Roles of
enzymatic and nonenzymatic antioxidants in plants during abiotic stress. Crit. Rev.
Biotechnol. 30, 161–175. doi: 10.3109/07388550903524243
Armengaud, J. (2013). Microbiology and proteomics, getting the best of both worlds!
Environ. Microbiol. 15, 12–23. doi: 10.1111/j.1462-2920.2012.02811.x
Armengaud, J., Christie-Oleza, J. A., Clair, G., Malard, V., and Duport, C. (2012).
Exoproteomics: exploring the world around biological systems. Expert Rev. Proteomics
9, 561–575. doi: 10.1586/epr.12.52
Babalola, O. O., and Glick, B. R. (2012). e use of microbial inoculants in African
agriculture: current practice and future prospects. J. Food Agric. Environ. 10, 540–549.
doi: 10.5897/SRE11.1714
Bartels, F. W., Baumgarth, B., Anselmetti, D., Ros, R., and Becker, A. (2003). Specic
binding of the regulatory protein ExpG to promoter regions of the galactoglucan biosynthesis
gene cluster of Sinorhizobium meliloti--a combined molecular biology and force spectroscopy
investigation. J. Struct. Biol. 143, 145–152. doi: 10.1016/S1047-8477(03)00127-8
Bashan, Y., De-Bashan, L. E., Prabhu, S. R., and Hernandez, J. (2014). Advances
in plant growth-promoting bacterial inoculant technology: formulations and
practical perspectives (1998–2013). Plant Soil 378, 1–33. doi: 10.1007/
s11104-013-1956-x
Becker, A., Rüberg, S., Küster, H., Roxlau, A. A., Keller, M., Ivashina, T., et al. (1997).
e 32-kilobase exp gene cluster of Rhizobium meliloti directing the biosynthesis of
galactoglucan: genetic organization and properties of the encoded gene products. J.
Bacteriol. 179, 1375–1384. doi: 10.1128/jb.179.4.1375-1384.1997
Bernhard, E. (2012). e bacterial phosphoenolpyruvate: sugar phosphotransferase
system (PTS): an interface between energy and signal transduction. J. Iran. Chem. Soc.
10, 593–630. doi: 10.1007/s13738-012-0185-1
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 10 frontiersin.org
Bojanovic, K., D'Arrigo, I., and Long, K. S. (2017). Global transcriptional responses
to osmotic, oxidative, and imipenem stress conditions in Pseudomonas putida. Appl.
Environ. Microbiol. 83, e03236–e03216. doi: 10.1128/AEM.03236-16
Borisova, M., Gaupp, R., Duckworth, A., Schneider, A., Dalügge, D., Mühleck, M.,
et al. (2016). Peptidoglycan recycling in gram-positive bacteria is crucial for survival in
stationary phase. MBio 7, e00923–e00916. doi: 10.1128/mBio.00923-16
Cappellari, L. R., and Banchio, E. (2020). Microbial volatile organic compounds
produced by Bacillus amyloliquefaciens GB03 ameliorate the eects of salt stress in
Mentha piperita principally through acetoin emission. J. Plant Growth Reg. 39, 764–775.
doi: 10.1007/s00344-019-10020-3
Cappellari, L. D. R., Chiappero, J., Palermo, T. B., Giordano, W., and Banchio, E.
(2020). Volatile organic compounds from rhizobacteria increase the biosynthesis of
secondary metabolites and improve the antioxidant status in Mentha piperita L. Grown
under salt stress. Agronomy 10:1094. doi: 10.3390/agronomy10081094
Chacko, S., Ramteke, P. W., and John, S. A. (2009). Amidase from plant growth
promoting rhizobacterium. J. Bacteriol. Res. 1, 46–50.
Chae, H. Z., Chung, S. J., and Rhee, S. G. (1994). ioredoxin-dependent peroxide
reductase from yeast. J. Biol. Chem. 269, 27670–27678. doi: 10.1016/
S0021-9258(18)47038-X
Cheng, C., Dong, Z., Han, X., Wang, H., Jiang, L., Sun, J., et al. (2017). ioredoxin
a is essential for motility and contributes to host infection of Listeria monocytogenes
via redox interactions. Front. Cell. Infect. Microbiol. 7:287. doi: 10.3389/
fcimb.2017.00287
Choby, J. E., and Skaar, E. P. (2016). Heme synthesis and acquisition in bacterial
pathogens. J. Mol. Biol. 28 428, 3408–3428. doi: 10.1016/j.jmb.2016.03.018
Clauss, K., von Roepenack-Lahaye, E., Bottcher, C., Roth, M. R., Welti, R., Erban, A.,
et al. (2011). Overexpression of sinapine esterase BnSCE3in oilseed rape seeds triggers
global changes in seed metabolism. Plant Physiol. 155, 1127–1145. doi: 10.1104/
pp.110.169821
Coba de la Peña, T., Redondo, F. J., Fillat, M. F., Lucas, M. M., and Pueyo, J. J. (2013).
Flavodoxin overexpression confers tolerance to oxidative stress in benecial soil bacteria
and improves survival in the presence of the herbicides paraquat and atrazine. J. Appl.
Microbiol. 115, 236–246. doi: 10.1111/jam.12224
Daneshmand, F., and Oloumi, H. (2015). e exogenously applied 5-Aminolevulinic
acid (ALA) mitigates salt stress in tomato plants. J. Crop. Prod. Process. 5, 135–148. doi:
10.18869/acadpub.jcpp.5.17.135
Das, K. C., and Das, C. K. (2000). ioredoxin, a singlet oxygen quencher and
hydroxyl radical scavenger: redox independent functions. Biochem. Biophys. Res.
Commun. 277, 443–447. doi: 10.1006/bbrc.2000.3689
Deochand, D. K., and Grove, A. (2017). MarR family transcription factors: dynamic
variations on a common scaold. Crit. Rev. Biochem. Mol. Biol. 52, 595–613. doi:
10.1080/10409238.2017.1344612
Desvaux, M., Dumas, E., Chafsey, I., Chambon, C., and Hébraud, M. (2010).
Comprehensive appraisal of the extracellular proteins from a monoderm bacterium:
theoretical and empirical exoproteomes of Listeria monocytogenes EGD-e by
Secretomics. J. Proteome Res. 9, 5076–5092. doi: 10.1021/pr1003642
Dolui, A. K., and Vijayaraj, P. (2020). Functional omics identies serine hydrolases
that mobilize storage lipids during rice seed germination. Plant Physiol. 184, 693–708.
doi: 10.1104/pp.20.00268
Dos Santos, C. V., and Rey, P. (2006). Plant thioredoxins are key actors in the oxidative
stress response. Trends Plant Sci. 11, 329–334. doi: 10.1016/j.tplants.2006.05.005
Du, Y., Shi, W. W., He, Y. X., Yang, Y. H., Zhou, C. Z., and Chen, Y. (2011). Structures
of the substrate-binding protein provide insights into the multiple compatible solute
binding specicities of the Bacillus subtilis ABC transporter OpuC. Biochem. J. 436,
283–289. doi: 10.1042/BJ20102097
Duan, Y., Chen, R., Zhang, R., Jiang, W., Chen, X., Yin, C., et al. (2021). Isolation,
identication, and antibacterial mechanisms of Bacillus amyloliquefaciens QSB-6 and its
eect on plant roots. Front. Microbiol. 12:746799. doi: 10.3389/fmicb.2021.746799
Engelhardt, E. (2007). Are S-layers exoskeletons? e basic function of protein surface
layers revisited. J. Struct. Biol. 160, 115–124. doi: 10.1016/j.jsb.2007.08.003
Gale, R. T., Li, F. K. K., Sun, T., Strynadka, N. C. J., and Brown, E. D. (2017). B. subtilis
LytR-CpsA-Psr enzymes transfer wall teichoic acids from authentic lipid-linked
substrates to mature peptidoglycan invitro. Cell. Chem. Biol. 24, 1537–1546.e4. doi:
10.1016/j.chembiol.2017.09.006
Galperin, M. Y. (2010). Diversity of structure and function of response regulator
output domains. Curr. Opin. Microbiol. 13, 150–159. doi: 10.1016/j.mib.2010.01.005
Gao, R., Mack, T. R., and Stock, A. M. (2007). Bacterial response regulators: versatile
regulatory strategies from common domains. Trends Biochem. Sci. 32, 225–234. doi:
10.1016/j.tibs.2007.03.002
García-García, A. L., García-Machado, F. J., Borges, A. A., Morales-Sierra, S., Boto, A.,
and Jiménez-Arias, D. (2020). Pure organic active compounds against abiotic stress: a
biostimulant overview. Front. Plant Sci. 11:575829. doi: 10.3389/fpls.2020.575829
Geigenberger, P., ormählen, I., Daloso, D. M., and Fernie, A. R. (2017). e
unprecedented versatility of the plant thioredoxin system. Trends Plant Sci. 22, 249–262.
doi: 10.1016/j.tplants.2016.12.008
Genişel, M., and Erdal, S. (2016). Alleviation of salt-induced oxidative damage by
5-aminolevulinic acid in wheat seedlings. AIP Conf. Proc. 1726:020025. doi:
10.1063/1.4945851
Ghosh, A., Manton, J. D., Mustafa, A. R., Gupta, M., Ayuso-Garcia, A., Rees, E. J., et al.
(2018). Proteins encoded by the gerP operon are localized to the inner coat in Bacillus
cereus spores and are dependent on GerPA and SafA for assembly. Appl. Environ.
Microbiol. 84, e00760–e00718. doi: 10.1128/AEM.00760-18
Gray, E. J., Di Falco, M. R., Souleimanov, A., and Smith, D. L. (2006b). Proteomic
analysis of the bacteriocin thuricin 17 produced by Bacillus thuringiens is NEB17. FEMS
Microbiol. Lett. 255, 27–32. doi: 10.1111/j.1574-6968.2005.00054.x
Gray, E. J., Lee, K. D., Souleimanov, A. M., Di Falco, M. R., Zhou, X., Ly, A., et al.
(2006a). A novel bacteriocin, thuricin 17, produced by plant growth promoting
rhizobacteria strain Bacillus thuringiensis NEB17: isolation and classication. J. Appl.
Microbiol. 100, 545–554. doi: 10.1111/j.1365-2672.2006.02822.x
Grosu-Tudor, S. S., Brown, L., Hebert, E. M., Brezeanu, A., Brinzan, A., Fadda, S., et al.
(2016). S-layer production by Lactobacillus acidophilu s IBB 801 under environmental stress
conditions. Appl. Microbiol. Biotechnol. 100, 4573–4583. doi: 10.1007/s00253-016-7355-5
Grove, A. (2013). MarR family transcription factors. Curr. Biol. 23, R142–R143. doi:
10.1016/j.cub.2013.01.013
Hachicho, N., Birnbaum, A., and Heipieper, H. J. (2017). Osmotic stress in colony and
planktonic cells of Pseudomonas putida mt-2 revealed signicant dierences in adaptive
response mechanisms. AMB Expr. 7:62. doi: 10.1186/s13568-017-0371-8
Hederstedt, L. (2012). Heme a biosynthesis. Biochim. Biophys. Acta 1817, 920–927.
doi: 10.1016/j.bbabio.2012.03.025
Hu, Y., Hu, Q., Wei, R., Li, R., Zhao, D., Ge, M., et al. (2019). e XRE family
transcriptional regulator SrtR in Streptococcus suis is involved in oxidant tolerance and
virulence. Front. Cell. Infect. Microbiol. 8:452. doi: 10.3389/fcimb.2018.00452
Hui, L., Lang, K., and Liang-Ju, W. (2006). Promotion of 5-aminolevunlinic acid on
seed germination of watermelon (Citrullus lanatus)under salt stress. J. Fruit Sci. 23,
854–859.
Hynönen, U., and Palva, A. (2013). Lactobacillus surface layer proteins: structure,
function, and applications. Appl. Microbiol. Biotechnol. 97, 5225–5243. doi: 10.1007/
s00253-013-4962-2
Jahreis, K., Pimentel-Schmitt, E. F., Brückner, R., and Titgemeyer, F. (2008). Ins and
outs of glucose transport systems in eubacteria. FEMS Microbiol. Rev. 32, 891–907. doi:
10.1111/j.1574-6976.2008.00125.x
Janczarek, M., Vinardell, J. M., Lipa, P., and Karaś, M. (2018). Hanks-type serine/
threonine protein kinases and phosphatases in bacteria: roles in Signaling and
adaptation to various environments. Int. J. Mol. Sci. 19:2872. doi: 10.3390/ijms19102872
Jason, M., Elena, L., and Ming, Z. (2014). Structural insight into the PTS sugar
transporter EIIC. Biochim. Biophys. Acta 1850. doi: 10.1016/j.bbagen.2014.03.013
Jedelská, T., Luhová, L., and Petřivalský, M. (2020). ioredoxins: emerging players in
the regulation of protein S-Nitrosation in plants. Plants (Basel) 9:1426. doi: 10.3390/
plants9111426
Karpievitch, Y. V., Ashoka, D. P., Gordon, A. A., Richard, D. S., and Alan, R. D. (2010).
Liquid chromatography mass spectrometry-based proteomics: biological and
technological aspects. Ann. Appl. Stat., 1797–1823. doi: 10.1214/10-AOAS341
Kawai, Y., Marles-Wright, J., Cleverley, R. M., Emmins, R., Ishikawa, S., Kuwano, M.,
et al. (2011). A widespread family of bacterial cell wall assembly proteins. EMB J. 30,
4931–4941. doi: 10.1038/emboj.2011.358
Kazerooni, E. A., Maharachchikumbura, S. S. N., Adhikari, A., Al-Sadi, A. M.,
Kang, S. M., Kim, L. R., et al. (2021). Rhizospheric Bacillus amyloliquefaciens protects
Capsicum annuum cv. Geumsugangsan from multiple abiotic stresses via multifarious
plant growth-promoting attributes. Front. Plant Sci. 12:669693. doi: 10.3389/
fpls.2021.669693
Keller, A., Nesvizhskii, A. I., Kolker, E., and Aebersold, R. (2002). Empirical statistical
model to estimate the accuracy of peptide identications made by MS/MS and data base
search. Anal. Chem. 74, 5383–5392. doi: 10.1021/ac025747h
Klotz, C., Goh, Y. J., O’Flaherty, S., Johnson, B., and Barrangou, R. (2020). Deletion of
S-layer associated Ig-like domain protein disrupts the Lactobacillus acidophilus cell
surface. Front. Microbiol. 11:345. doi: 10.3389/fmicb.2020.00345
Krämer, R. (2010). Bacterial stimulus perception and signal transduction: response to
osmotic stress. Chem. Rec. 10, 217–229. doi: 10.1002/tcr.201000005
Kucharova, V., and Wiker, H. G. (2014). Proteogenomics in microbiology: taking the
right turn at the junction of genomics and proteomics. Proteomics 14, 2660–2675. doi:
10.1002/pmic.201400168
Lee, J., Sands, Z. A., and Biggin, P. C. (2016). A Numbering System for MFS
Transporter Proteins. Front. Mol. Biosci. 3:21. doi: 10.3389/fmolb.2016.00021
Lewin, A., and Hederstedt, L. (2016). Heme a synthase in bacteria depends on one
pair of cysteinyls for activity. Biochim. Biophys. Acta 1857, 160–168. doi: 10.1016/j.
bbabio.2015.11.008
Li, Y. C., Ren, J. P., Cho, M. J., Zhou, S. M., Kim, Y. B., Guo, H. X., et al. (2009). e
level of expression of thioredoxin is linked to fundamental properties and applications
of wheat seeds. Mol. Plant 2, 430–441. doi: 10.1093/mp/ssp025
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 11 frontiersin.org
Listgarten, J., and Emili, A. (2005). Statistical and computational methods for
comparative proteomic profiling using liquid chromatography-tandem mass
spectrometry. Mol. Cell. Proteomics 4, 419–434. doi: 10.1074/mcp.R500005-
MCP200
Liu, Z., Zhang, D., Ning, F., Zhang, S., Hou, Y., Gao, M., et al. (2023). Resistance and
adaptation of mature algal-bacterial granular sludge under salinity stress. Sci. Total
Environ. 861:160558. doi: 10.1016/j.scitotenv.2022.160558
Lomelino, C. L., Andring, J. T., McKenna, R., and Kilberg, M. S. (2017). Asparagine
synthetase: function, structure, and role in disease. J. Biol. Chem. 292, 19952–19958. doi:
10.1074/jbc.R117.819060
Lowry, O. H., Rosebrough, N. J., Farr, A. L., and Randall, R. J. (1951). Protein
measurement with the Folin phenol reagent. J. Biol. Chem. 193, 265–275. doi: 10.1016/
S0021-9258(19)52451-6
Lu, J., and Holmgren, A. (2014). e thioredoxin antioxidant system. Free Radic. Biol.
Med. 66, 75–87. doi: 10.1016/j.freeradbiomed.2013.07.036
Lycklama, A., Nijeholt, J. A., Vietrov, R., Schuurman-Wolters, G. K., and Poolman, B.
(2018). Energy coupling eciency in the type IABC transporter GlnPQ. J. Mol. Biol.
430, 853–866. doi: 10.1016/j.jmb.2018.02.001
Ma, J., Lei, H. T., Reyes, F. E., Sanchez-Martinez, S., Sarhan, M. F., Hattne, J., et al.
(2019). Structural basis for substrate binding and specicity of a sodium-alanine
symporter AgcS. Proc. Natl. Acad. Sci. U. S. A. 116, 2086–2090. doi: 10.1073/
pnas.1806206116
Maeda, K., Finnie, C., Østergaard, O., and Svensson, B. (2003). Identication, cloning
and characterization of two thioredoxin h isoforms, HvTrxh1 and HvTrxh2, from the
barley seed proteome. Eur. J. Biochem. 270, 2633–2643. doi:
10.1046/j.1432-1033.2003.03637.x
Maeda, K., Finnie, C., and Svensson, B. (2005). Identication of thioredoxin
h-reducible disulphides in proteomes by dierential labelling of cysteines: insight into
recognition and regulation of proteins in barley seeds by thioredoxin h. Proteomics 5,
1634–1644. doi: 10.1002/pmic.200401050
Martínez, J. L., Petranovic, D., and Nielsen, J. (2016). Heme metabolism in stress
regulation and protein production: from Cinderella to a key player. Bioengineered 7,
112–115. doi: 10.1080/21655979.2015.1126016
Meng, L., Wong, J. H., Feldman, L. J., Lemaux, P. G., and Buchanan, B. B. (2010). A
membrane-associated thioredoxin required for plant growth moves from cell to cell,
suggestive of a role in intercellular communication. Proc. Natl. Acad. Sci. U. S. A. 107,
3900–3905. doi: 10.1073/pnas.0913759107
Meyer, Y., Belin, C., Delorme-Hinoux, V., Reichheld, J.-P., and Riondet, C. (2012).
ioredoxin and Glutaredoxin Systems in Plants: molecular mechanisms, Crosstalks,
and functional signicance. Antioxid. Redox Signal. 17, 1124–1160. doi: 10.1089/
ars.2011.4327
Molloy, S. (2011). LCP proteins take the nal step. Nat. Rev. Microbiol. 9:768. doi:
10.1038/nrmicro2684
Moy, B. E., and Seshu, J. (2021). STAS domain only proteins in bacterial gene
regulation. Front. Cell. Infect. Microbiol. 11:679982. doi: 10.3389/fcimb.2021.679982
Msimbira, L. A., Subramanian, S., Naamala, J., Antar, M., and Smith, D. L. (2022).
Secretome analysis of the plant biostimulant bacteria strains Bacillus subtilis (EB2004S)
and Lactobacillus helveticus (EL2006H) in response to pH changes. Int. J. Mol. Sci.
23:15144. doi: 10.3390/ijms232315144
Mueller, E. A., and Levin, P. A. (2020). Bacterial Cell Wall quality control during
environmental stress. MBio 11, e02456–e02420. doi: 10.1128/mBio.02456-20
Naamala, J., Msimbira, L. A., Antar, M., Subramanian, S., and Smith, D. L. (2022).
Cell-free supernatant obtained from a salt tolerant Bacillus amyloliquefaciens strain
enhances germination and radicle length under NaCl stressed and optimal conditions.
Front. Sustain. Food Syst. 6:788939. doi: 10.3389/fsufs.2022.788939
Naamala, J., Msimbira, L. A., Subramanian, S., and Smith, D. L. (2023). Lactobacillus
helveticus EL2006H cell-free supernatant enhances growth variables in Zea mays
(maize), Glycine max L. Merill (soybean) and Solanum tuberosum (potato) exposed to
NaCl stress. Front. Microbiol. 13:1075633. doi: 10.3389/fmicb.2022.1075633
Naamala, J., and Smith, D. L. (2020). Relevance of plant growth promoting
microorganisms and their derived compounds, in the face of climate change. Agronomy
10:1179. doi: 10.3390/agronomy10081179
Nadeem, S. J., Ahmad, M., Zahir, Z. A., Javaid, A., and Ashraf, M. (2015). e role of
mycorrhizae and plant growth promoting rhizobacteria (PGPR) in improving crop
productivity under stressful environments. Biotechnol. Adv. 32, 429–448. doi: 10.1016/j.
biotechadv.2013.12.005
Ngalimat, M. S., Yahaya, R. S. R ., Baharudin, M. M. A., Yaminudin, S. M., Karim, M.,
Ahmad, S. A., et al. (2021). A review on the biotechnological applications of the
operational group Bacillus amyloliquefaciens. Microorganisms 9:614. doi: 10.3390/
microorganisms9030614
Niklas, K. J., Cobb, E. D., and Matas, A. J. (2017). e evolution of hydrophobic cell
wall biopolymers: from algae to angiosperms. J. Exp. Bot. 68, 5261–5269. doi: 10.1093/
jxb/erx215
Ofengand, J. (2002). Ribosomal RNA pseudouridines and pseudouridine synthases.
FEBS Lett. 514, 17–25. doi: 10.1016/S0014-5793(02)02305-0
Oren, A. (2002). “Adaptation of halophilic archaea to life at high salt concentrations”
in Salinity: Environment plants molecules. eds. A. Läuchli and U. Lüttge (Dordrecht:
Springer)
Oren, A. (2008). Microbial life at high salt concentrations: phylogenetic and metabolic
diversity. Saline Syst. 4:2. doi: 10.1186/1746-1448-4-2
Palomino, M. M., Sanchez-Rivas, C., and Ruzal, S. M. (2009). High salt stress in
Bacillus subtilis: involvement of PBP4* as a peptidoglycan hydrolase. Res. Microbiol. 160,
117–124. doi: 10.1016/j.resmic.2008.10.011
Park, J. T. (1995). Why does Escher ichia coli recycle its cell wall peptides? Mol.
Microbiol. 17, 421–426. doi: 10.1111/j.1365-2958.1995.mmi_17030421.x
Pasqua, M., Grossi, M., Zennaro, A., Fanelli, G., Micheli, G., Barras, F., et al. (2019).
e varied role of eux pumps of the MFS family in the interplay of bacteria with
animal and plant cells. Microorganisms, 22 7:285. doi: 10.3390/microorganisms7090285
Pérez-Rueda, E., Collado-Vides, J., and Segovia, L. (2004). Phylogenetic distribution
of DNA-binding transcription factors in bacteria and archaea. Comput. Biol. Chem. 28,
341–350. doi: 10.1016/j.compbiolchem.2004.09.004
Philippe, G., Sorensen, I., Jiao, C., Sun, X., Fei, Z., Domozych, D. S., et al. (2020). Cutin
and suberin: assembly and origins of specialized lipidic cell wall scaolds. Curr. Opin.
Plant Biol. 55, 11–20. doi: 10.1016/j.pbi.2020.01.008
Piechulla, B., Lemfack, M. C., and Kai, M. (2017). Eects of discrete bioactive
microbial volatiles on plants and fungi. Plant Cell Environ. 40, 2042–2067. doi: 10.1111/
pce.13011
Pittman, J. R., Buntyn, J. O., Posadas, G., Nanduri, B., Pendarvis, K., and
Donaldson, J. R. (2014). Proteomic analysis of cross protection provided between cold
and osmotic stress in Listeria monocytogenes. J. Proteome Res.4 13, 1896–1904. doi:
10.1021/pr401004a
Popham, D. L., and Young, K. D. (2003). Role of penicillin-binding proteins in
bacterial cell morphogenesis. Curr. Opin. Microbiol. 6, 594–599. doi: 10.1016/j.
mib.2003.10.002
Prithiviraj, B., Zhou, X., Souleimanov, A., Khan, W. M., and Smith, D. L. (2003). A
host-specic bacteria-to-plant signal molecule (nod factor) enhances germination and
early growth of diverse crop plants. Planta 216, 437–445. doi: 10.1007/s00425-002-0928-9
Pumirat, P., Saetun, P., Sinchaikul, S., Chen, S. T., Korbsrisate, S., and
ongboonkerd, V. (2009). Altered secretome of Burkholderia pseudomallei induced by
salt stress. Biochim. Biophys. Acta 1794, 898–904. doi: 10.1016/j.bbapap.2009.01.011
Rhazi, L., Cazalis, R., and Aussenac, T. (2003). Sulydryl-disulde changes in storage
proteins of developing wheat grain: inuence on the SDS-unextractable glutenin
polymer formation. J. Cereal Sci. 38, 3–13. doi: 10.1016/S0733-5210(03)00019-5
Roovers, M., Kaminska, K. H., Tkaczuk, K. L., Gigot, D., Droogmans, L., and
Bujnicki, J. M. (2008). The YqfN protein of Bacillus subtilis is the tRNA: m1A22
methyltransferase (TrmK). Nucleic Acids Res. 36, 3252–3262. doi: 10.1093/nar/
gkn169
Rubiano-Labrador, C., Bland, C., Miotello, G., Armengaud, J., and Baena, S. (2015).
Salt stress induced changes in the exoproteome of the halotolerant bacterium Tistli a
consotensis deciphered by Proteogenomics. PLoS One 10:e0135065. doi: 10.1371/journal.
pone.0135065
Sauvage, E., Ker, F., Terrak, M., Ayala, J. A., and Charlier, P. (2008). e penicillin-
binding proteins: structure and role in peptidoglycan biosynthesis. FEMS Microbiol. Rev.
32, 234–258. doi: 10.1111/j.1574-6976.2008.00105.x
Schoof, M., O’Callaghan, M., Sheen, C. R., Glare, T. R., and Hurst, M. R. H. (2022).
Identication of genes involved in exoprotein release using a high throughput
exoproteome screening assay in Yersinia entomophaga. PLoS One 17:e0263019. doi:
10.1371/journal.pone.0263019
Schwinghamer, T., Souleimanov, A., Dutilleul, P., and Smith, D. (2016). The
response of canola cultivars to lipochitooligosaccharide (nod Bj V [C18:1, MeFuc])
and thuricin 17. Plant Growth Regul. 78, 421–434. doi: 10.1007/s10725-015-0104-4
Sengupta, R., Altermann, E., Anderson, R. C., McNabb, W. C., Moughan, P. J., and
Roy, N. C. (2013). e role of cell surface architecture of lactobacilli in host-microbe
interactions in the gastrointestinal tract. Mediators Inamm. 2013:237921. doi:
10.1155/2013/237921
Senzani, S., Li, D., Bhaskar, A., Ealand, C., Chang, J., Rimal, B., et al. (2017). An
amidase_3 domain-containing N-acetylmuramyl-L-alanine amidase is required for
mycobacterial cell division. Sci. Rep. 7:1140. doi: 10.1038/s41598-017-
01184-7
Setlow, P. (2014). Germination of spores of bacillus species: what weknow and do not
kn ow. J. Bacteriol. 196, 1297–1305. doi: 10.1128/JB.01455-13
Sharma, A. K., Rigby, A. C., and Alper, S. L. (2011). STAS domain structure and
function. Cell. Physiol. Biochem. 28, 407–422. doi: 10.1159/000335104
Shen, G., Sun, W., Chen, Z., Shi, L., Hong, J., and Shi, J. (2022). Plant GDSL Esterases/
lipases: evolutionary, physiological and molecular functions in plant development.
Plants (Basel) 11, –468. doi: 10.3390/plants11040468
Shin, J.-H., Sulpizio, A. G., Kelley, A., Alvarez, L., Murphy, S. G., Fan, L., et al. (2020).
Structural basis of peptidoglycan endopeptidase regulation. Proc. Natl. Acad. Sci. 117,
11692–11702. doi: 10.1073/pnas.2001661117
Naamala et al. 10.3389/fmicb.2023.1206152
Frontiers in Microbiology 12 frontiersin.org
Si, M., Chen, C., Zhong, J., Li, X., Liu, Y., Su, T., et al. (2020). MsrR is a thiol-based
oxidation-sensing regulator of the XRE family that modulates C. glutamicum oxidative
stress resistance. Microb. Cell Factories 19:189. doi: 10.1186/s12934-020-01444-8
Siegel, S. D., Amer, B. R., Wu, C., Sawaya, M. R., Gosschalk, J. E., Clubb, R. T., et al.
(2019). Structure and mechanism of LcpA, a phosphotransferase at mediates
glycosylation of a gram-positive bacterial Cell Wall-anchored protein. MBio 10, e01580–
e01518. doi: 10.1128/mBio.01580-18
Singleton, P. W., El Swaify, S. A., and Bohlool, B. B. (1982). Eect of salinity on
rhizobium growth and survival. Appl. Environ. Microbiol. 44, 884–890. doi: 10.1128/
aem.44.4.884-890.1982
Soussi, M., Santamaría, M., Ocaña, A., and Lluch, C. (2001). Eects of salinity on
protein and lipopolysaccharide pattern in a salt-tolerant strain of Mesorhizobium ciceri.
J. Appl. Microbiol. 90, 476–481. doi: 10.1046/j.1365-2672.2001.01269.x
Spaepen, S., Vanderleyden, J., and Remans, R. (2007). Indole-3-acetic acid in microbial
and microorganism-plant signaling. FEMS Microbiol. Rev. 31, 425–448. doi:
10.1111/j.1574-6976.2007.00072.x
Subramanian, S., Souleimanov, A., and Smith, D. L. (2021). Thuricin17
production and proteome differences in Bacillus thuring iensis NEB17 cell-free
supernatant under NaCl stress. Front. Sustain. Food Syst. 5:630628. doi: 10.3389/
fsufs.2021.630628
Sweeney, P., Galliford, A., Kumar, A., Raju, D., Krishna, N. B., Sutherland, E., et al.
(2022). Structure, dynamics, and molecular inhibition of the Staphylococcus aureus
m1A22-tRNA methyltransferase TrmK. J. Biol. Chem. 298:102040. doi: 10.1016/j.
jbc.2022.102040
Takahashi, K., Shimada, T., Kondo, M., Tamai, A., Mori, M., Nishimura, M., et al.
(2010). Ectopic expression of an esterase, which is a candidate for the unidentied plant
cutinase, causes cuticular defects in Arabidopsis thaliana. Plant Cell Physiol. 51, 123–131.
doi: 10.1093/pcp/pcp173
Teichmann, L., Kümmel, H., Warmbold, B., and Bremer, E. (2018). OpuF, a new
bacillus compatible solute ABC transporter with a substrate-binding protein fused to
the transmembrane domain. Appl. Environ. Microbiol. 84, e01728–e01718. doi: 10.1128/
AEM.01728-18
Tsuzuki, M., Moskvin, O. V., Kuribayashi, M., Sato, K., Retamal, S., Abo, M., et al.
(2011). Salt stress-induced changes in the transcriptome, compatible solutes, and
membrane lipids in the facultatively phototrophic bacterium Rhodobacter
sphaeroides. Applied Environ Microbiol. 77, 7551–7559. doi: 10.1128/AEM.
05463-11
Unkefer, P. J., Knight, T. J., and Martinez, R. A. (2023). e intermediate in a nitrate-
responsive ω-amidase pathway in plants may signal ammonium assimilation status.
Plant Physiol. 191, 715–728. doi: 10.1093/plphys/kiac501
Ursache, R., De Jesus, V. T. C., Denervaud, T. V., Gully, K., De Bellis, D.,
Schmid-Siegert, E., et al. (2021). GDSL-domain proteins have key roles in suberin
polymerization and degradation. Nat. Plants. 7, 353–364. doi: 10.1038/
s41477-021-00862-9
Weber, B. W., Kimani, S. W., Varsani, A., Cowan, D. A., Hunter, R., Venter, G. A., et al.
(2013). e mechanism of the amidases: mutating the glutamate adjacent to the catalytic
triad inactivates the enzyme due to substrate mispositioning. J. Biol. Chem. 288,
28514–28523. doi: 10.1074/jbc.M113.503284
Will, W. R., and Fang, F. C. (2020). e evolution of MarR family transcription factors
as counter-silencers in regulatory networks. Curr. Opin. Microbiol. 55, 1–8. doi:
10.1016/j.mib.2020.01.002
Woldemariam, Y. K., Wan, Z., Yu, Q., Li, H., Wei, X., Liu, Y., et al. (2020). Prebiotic,
probiotic, antimicrobial, and functional food applications of Bacillus amyloliquefaciens.
J. Agric. Food Chem. 68, 14709–14727. doi: 10.1021/acs.jafc.0c06396
Woodson, J. D., Perez-Ruiz, J. M., and Chory, J. (2011). Heme synthesis by plastid
ferrochelatase Iregulates nuclear gene expression in plants. Curr. Biol. 21, 897–903. doi:
10.1016/j.cub.2011.04.004
Woolley, R. C., Vediyappan, G., Anderson, M., Lackey, M., Ramasubramanian, B.,
Jiangping, B., et al. (2005). Characterization of the Vibrio cholerae vceCAB multiple-drug
resistance eux operon in Escherichia coli. J. Bacteriol. 187, 5500–5503. doi: 10.1128/
JB.187.15.5500-5503.2005
Wu, Y., Li, J., Wang, J., Dawuda, M. M., Liao, W., Meng, X., et al. (2022). Heme is
involved in the exogenous ALA-promoted growth and antioxidant defense system of
cucumber seedlings under salt stress. BMC Plant Biol. 22:329. doi: 10.1186/
s12870-022-03717-3
Xu, H., Gao, J., Portieles, R., Du, L., Gao, X., and Borras-Hidalgo, O. (2022).
Endophytic bacterium Bacillus aryabhattai induces novel transcriptomic changes
to stimulate plant growth. PLoS One 17:e0272500. doi: 10.1371/journal.
pone.0272500
Yan, N. (2013). Structural advances for the major facilitator superfamily (MFS)
transporters. Trends Biochem. Sci. 38, 151–159. doi: 10.1016/j.tibs.2013.01.003
Yan, N., Marschner, P., Cao, W., Zuo, C., and Qin, W. (2015). Inuence of salinity and
water content on soil microorganisms. Int. Soil Water Conserv. Res. 3, 316–323. doi:
10.1016/j.iswcr.2015.11.003
Yin, J., Ren, W., Huang, X., Deng, J., Li, T., and Yin, Y. (2018). Potential mechanisms
connecting purine metabolism and cancer therapy. Front. Immunol. 9:1697. doi:
10.3389/mmu.2018.01697
Zahid, A., Afoulous, S., and Cazalis, R. (2008). ioredoxin h system and wheat seed
quality. Cereal Chem. 85, 799–807. doi: 10.1094/CCHEM-85-6-0799
Zahran, H. H. (1997). Diversity, adaptation, and activity of the bacterial ora in saline
environments. Biol. Fertil. Soils 25, 211–223. doi: 10.1007/s003740050306
Zeller, T., and Klug, G. (2006). ioredoxins in bacteria: functions in oxidative stress
response and regulation of thioredoxin genes. Naturwissenschaen 93, 259–266. doi:
10.1007/s00114-006-0106-1
Zhang, X., Fang, A., Riley, C. P., Wang, M., Regnier, F. E., and Buck, C. (2010). Multi-
dimensional liquid chromatography in proteomics–a review. Anal. Chim. Acta 664,
101–113. doi: 10.1016/j.aca.2010.02.001
Zhang, Z. W., Feng, L. Y., Cheng, J., Tang, H., Xu, F., Zhu, F., et al. (2013). e roles of
two transcription factors, ABI4 and CBFA, in ABA and plastid signalling and stress
responses. Plant Mol. Biol. 83, 445–458. doi: 10.1007/s11103-013-0102-8
Zhang, Y., Liang, S., Pan, Z., Yu, Y., Yao, H., Liu, Y., et al. (2022). XRE family
transcriptional regulator XtrSs modulates Streptococcus suis tness under hydrogen
peroxide stress. Arch. Microbiol. 204:244. doi: 10.1007/s00203-022-02854-5
Zhang, H. H., Wang, M. L., Li, Y. Q., Yan, W., Chang, Z. Y., Ni, H. L., et al. (2020).
GDSL esterase/lipases OsGELP34 and OsGELP110/OsGELP115 are essential for rice
pollen development. J. Integr. Plant Biol. 62, 1574–1593. doi: 10.1111/jipb.12919
Zhang, X. C., Zhao, Y., Heng, J., and Jiang, D. (2015). Energy coupling mechanisms of
MFS transporters. Protein Sci. 24, 1560–1579. doi: 10.1002/pro.2759
Zhao, Y., Dunker, W., Yu, Y. T., and Karijolich, J. (2018). e role of noncoding RNA
Pseudouridylation in nuclear gene expression events. Front. Bioeng. Biotechnol. 6:8. doi:
10.3389/ioe.2018.00008
Zhu, W., R adadiya, A., Bisson, C., Wenzel, S., Nordin, B. E., Martínez-Márquez, F., et al.
(2019). High-resolution crystal structure of human asparagine synthetase enables analysis
of inhibitor binding and selectivity. C ommun Biol 2:345. doi: 10.1038/s42003-019-0587-z
Zhu, H., Zhou, J., Ni, Y., Yu, Z., Mao, A., Hu, Y., et al. (2014). Contribution of
eukaryotic-type serine/threonine kinase to stress response and virulence of Streptococcus
suis. PLoS One 9:e91971. doi: 10.1371/journal.pone.0091971
Article
Full-text available
Plant growth promoting microorganisms and their derived compounds, such as cell-free supernatant (CFS), enhance plant growth under stressed and non stressed conditions. Such technology is sustainable and environmentally friendly, which is desirable amidst the climate change threat. The current study evaluated the effect of CFS obtained from Lactobacillus helveticus EL2006H on its ability to enhance mean percentage germination and mean radicle length of corn and soybean, as well as growth parameters of potato, using treatment formulations that consisted of 0.2 and 1.0% [v/v] L. helveticus EL2006H CFS concentrations and 100 mM NaCl and 150 mM NaCl levels. Results show that treatment with 100 mM NaCl lowered percentage germination of corn by 52.63%, at 72 h, and soybean by 50%, at 48 h. Treatment with 100 NaCl +0.2% EL2006H enhanced percentage germination of soybean by 44.37%, at 48 h, in comparison to that of the 100 mM NaCl control. One hundred mM NaCl lowered radicle length of corn and soybean by 38.58 and 36.43%, respectively. Treatment with 100 Mm NaCl +1.0% EL2006H significantly increased radicle length of corn by 23.04%. Treatment with 100 mM NaCl +0.2% EL2006H significantly increased photosynthetic rate, leaf greenness and fresh weight of potato. Increasing NaCl concentration to 150 NaCl lowered the effectiveness of the 0.2% EL2006H CFS on the same growth variables of potato. In general, the lower CFS concentration of 0.2% was more efficient at enhancing germination in soybean while the higher concentration of 1.0% was more efficient at enhancing radicle length of corn. There was an observed variation in the effectiveness of L. helveticus EL2006H CFS across the different CFS concentrations, NaCl levels and crop species studied. In conclusion, based on findings of this study, CFS obtained from L. helveticus can be used as a bio stimulant to enhance growth of corn, soybean and potato. However, further studies need to be conducted, for validation, especially under field conditions, for commercial application.
Article
Full-text available
It is well-known that there is a high frequency of plant-growth-promoting strains in Bacillus subtilis and that these can be effective under both stressful and stress-free conditions. There are very few studies of this activity in the case of Lactobacillus helveticus. In this study, the effects of pH on the secretome (proteins) in the cell-free supernatants of two bacterial strains were evaluated. The bacteria were cultured at pH 5, 7 and 8, and their secretome profiles were analyzed, with pH 7 (optimal growth pH) considered as the “control”. The results showed that acidity (lower pH 5) diminishes the detectable production of most of the secretome proteins, whereas alkalinity (higher pH 8) increases the detectable protein production. At pH 5, five (5) new proteins were produced by L. helveticus, including class A sortase, fucose-binding lectin II, MucBP-domain-containing protein, SLAP-domain-containing protein and hypothetical protein LHEJCM1006_11110, whereas for B. subtilis, four (4) types of proteins were uniquely produced (p ≤ 0.05), including helicase-exonuclease AddAB subunit AddB, 5-methyltetrahydropteroyltriglutamate-homocysteine S-methyltransferase, a cluster of ABC-F family ATP-binding-cassette-domain-containing proteins and a cluster of excinuclease ABC (subunit B). At pH 8, Bacillus subtilis produced 56 unique proteins. Many of the detected proteins were involved in metabolic processes, whereas the others had unknown functions. The unique and new proteins with known and unknown functions suggest potential the acclimatization of the microbes to pH stress.
Article
Full-text available
A metabolite of ammonium assimilation was previously theorized to be involved in the coordination of the overall nitrate response in plants. Here we show that 2-hydroxy-5-oxoproline, made by transamination of glutamine, the first product of ammonium assimilation, may be involved in signaling a plant's ammonium assimilation status. In leaves, 2-hydroxy-5-oxoproline met four foundational requirements to be such a signal. First, when it was applied to foliage, enzyme activities of nitrate reduction and ammonium assimilation increased; the activities of key tricarboxylic acid cycle-associated enzymes that help to supply carbon skeletons for amino acid synthesis also increased. Second, its leaf pools increased as nitrate availability increased. Third, the pool size of its precursor, Gln, reflected ammonium assimilation rather than photorespiration. Fourth, it was widely conserved among monocots, dicots, legumes and nonlegumes and in plants with C3 or C4 metabolism. Made directly from the first product of ammonium assimilation, 2-hydroxy-5-oxoproline acted as a nitrate uptake stimulant. When 2-hydroxy-5-oxoproline was provided to roots, the plant's nitrate uptake rate approximately doubled. Plants exogenously provided with 2-hydroxy-5-oxoproline to either roots or leaves accumulated greater biomass. A model was constructed that includes proposed roles of 2-hydroxy-5-oxoproline as a signal molecule of ammonium assimilation status in leaves, as a stimulator of nitrate uptake by roots and nitrate downloading from the xylem. In summary, a glutamine metabolite made in the ω-amidase pathway stimulated nitrate uptake by roots and was likely to be a signal of ammonium assimilation status in leaves. A chemical synthesis method for 2-hydroxy-5-oxoproline was also developed.
Article
Full-text available
In nature, plants interact with a wide range of microorganisms, and most of these microorganisms could induce growth through the activation of important molecular pathways. The current study evaluated whether the endophytic bacterium Bacillus aryabhattai encourages plant growth and the transcriptional changes that might be implicated in this effect. The endophytic bacterium promotes the growth of Arabidopsis and tobacco plants. The transcriptional changes in Arabidopsis plants treated with the bacterium were also identified, and the results showed that various genes, such as cinnamyl alcohol dehydrogenase, apyrase, thioredoxin H8, benzaldehyde dehydrogenase, indoleacetaldoxime dehydratase, berberine bridge enzyme-like and gibberellin-regulated protein, were highly expressed. Also, endophytic bacterial genes, such as arginine decarboxylase, D-hydantoinase, ATP synthase gamma chain and 2-hydroxyhexa-2,4-dienoate hydratase, were activated during the interaction. These findings demonstrate that the expression of novel plant growth-related genes is induced by interaction with the endophytic bacterium B . aryabhattai and that these changes may promote plant growth in sustainable agriculture.
Article
Full-text available
A biosynthetic precursor of tetrapyrrol, 5-aminolevulinic acid (ALA), is widely used in agricultural production, as an exogenous regulatory substance that effectively regulates plant growth. Previous studies have shown that heme and chlorophyll accumulate in plants under salt stress, when treated with exogenous ALA. In this study, we explored the regulatory role of heme in plants, by spraying 25 mg L –1 ALA onto the leaves of cucumber seedlings treated with heme synthesis inhibitor (2,2'-dipyridyl, DPD) and heme scavenger (hemopexin, Hx), under 50 mmol L –1 NaCl stress. The results showed that NaCl alone and DPD + Hx treatments to cucumber seedlings subjected to salt stress adversely affected their growth, by decreasing biomass accumulation, root activity, and root morphology. In addition, these treatments induced an increase in membrane lipid oxidation, as well as enhancement of anti-oxidase activities, proline content, and glutamate betaine. However, exogenous ALA application increased the plant growth and root architecture indices under NaCl stress, owing to a lack of heme in the seedlings. In addition, cucumber seedlings treated with DPD and Hx showed inhibition of growth under salt stress, but exogenous ALA effectively improved cucumber seedling growth as well as the physiological characteristics; moreover, the regulation of ALA in plants was weakened when heme synthesis was inhibited. Heme biosynthesis and metabolism genes, HEMH and HO1 , which are involved in the ALA metabolic pathway, were upregulated under salinity conditions, when ferrochelatase activity was inhibited. Application of exogenous ALA increased the heme content in the leaves. Thus, exogenous ALA may supplement the substrates for heme synthesis. These results indicated that heme plays a vital role in the response of plants to salinity stress. In conclusion, heme is involved in ALA-mediated alleviation of damage caused to cucumber seedlings and acts as a positive regulator of plant adaption.
Article
Full-text available
The enzyme m¹A22-tRNA methyltransferase (TrmK) catalyses the transfer of a methyl group to the N1 of adenine 22 in bacterial tRNAs. TrmK is essential for Staphylococcus aureus survival during infection, but has no homologue in mammals, making it a promising target for antibiotic development. Here we characterize the structure and function of S. aureus TrmK using X-ray crystallography, binding assays, and molecular dynamics simulations. We report crystal structures for the S. aureus TrmK apoenzyme as well as in complexes with methyl donor SAM and co-product product SAH. Isothermal titration calorimetry showed that SAM binds to the enzyme with favourable but modest enthalpic and entropic contributions, whereas SAH binding leads to an entropic penalty compensated for by a large favourable enthalpic contribution. Molecular dynamics simulations point to specific motions of the C-terminal domain being altered by SAM binding, which might have implications for tRNA recruitment. In addition, activity assays for S. aureus TrmK-catalysed methylation of A22 mutants of tRNALeu demonstrate that the adenine at position 22 is absolutely essential. In-silico screening of compounds suggested the multi-functional organic toxin plumbagin as a potential inhibitor of TrmK, which was confirmed by activity measurements. Furthermore, LC-MS data indicated the protein was covalently modified by one equivalent of the inhibitor, and proteolytic digestion coupled with LC-MS identified Cys92 in the vicinity of the SAM-binding site as the sole residue modified. These results identify a cryptic binding pocket of S. aureus TrmK, laying a foundation for future structure-based drug discovery.
Article
Full-text available
Streptococcus suis is an important emerging zoonosis that causes economic losses in the pig industry and severe threats to public health. Transcriptional regulators play essential roles in bacterial adaptation to host environments. In this study, we identified a novel XRE family transcriptional regulator in S. suis CZ130302, XtrSs, involved in the bacterial fitness to hydrogen peroxide stress. Based on electrophoretic mobility shift and β-galactosidase activity assays, we found that XtrSs auto-regulated its own transcription and repressed the expression of its downstream gene psePs, a surface protein with unknown function in S. suis, by binding to a palindromic sequence from the promoter region. Furthermore, we proved that the deletion of the psePs gene attenuated bacterial antioxidant response. Phylogenetic analysis revealed that XtrSs and PsePs naturally co-existed as a combination in most S. suis genomes. Collectively, we demonstrated the binding characteristics of XtrSs in S. suis and provided a new insight that XtrSs played a critical role in modulating psePs to the hydrogen peroxide resistance of S. suis.
Article
Full-text available
Seed germination and early plant growth are key stages in plant development that are, susceptible to salinity stress. Plant growth promoting microorganisms (PGPMs) produce substances, in their growth media, that could enhance plant growth under more optimal conditions, and or mitigate abiotic stresses, such as salinity. This study was carried out to elucidate the ability of a NaCl tolerant Bacillus amyloliquefaciens strain's cell-free supernatant to enhance germination and radicle length of corn and soybean, under optimal and NaCl stressed growth conditions. Three NaCl levels (0, 50, and 75 mM) and four cell-free supernatant concentrations (1.0, 0.2, 0.13, and 0.1% v/v) were used to formulate treatments that were used in the study. There were observed variations in the effect of treatments on mean radicle length and mean percentage germination of corn and soybean. Overall, the study showed that Bacillus amyloliquefaciens (BA) EB2003 cell-free supernatant could enhance mean percentage germination and or mean radicle length of corn and soybean. At optimal conditions (0 mM NaCl), 0.2% BA, 0.13% BA, and 0.1% BA concentrations resulted in 36.4, 39.70, and 39.91%, increase in mean radicle length of soybean, respectively. No significant observations were observed in mean radicle length of corn, and mean percentage germination of both corn and soybean. At 50 mM NaCl, 1.0% BA resulted in 48.65% increase in mean percentage germination of soybean, at 24 h. There was no observed effect of the cell-free supernatant on mean radicle length and mean percentage germination, at 72 and 48 h, in soybean. In corn however, at 50 mM NaCl, treatment with 0.2% BA and 0.13% BA enhanced mean radicle length by 23.73 and 37.5%, respectively. The resulting radicle lengths (43.675 and 49.7125 cm) were not significantly different from that of the 0 mM control. There was no observed significant effect of the cell-free supernatant on mean germination percentage of corn, at 50 mM NaCl. At 75 mM NaCl, none of the treatments enhanced mean radicle length or mean percentage germination to levels significantly higher than the 75 mM NaCl. Treatment with 1.0% BA, however, enhanced mean percentage germination to a level not significantly different from that of the 0 mM control, at 72 h. Likewise, in corn, none of the treatments enhanced radicle length to lengths significantly higher than the 75 mM control, although treatment with 1.0% BA, 0.13% BA, and 0.1% BA elongated radicles to lengths not significantly different from the 0 mM NaCl control. Treatment with 0.2% BA, 0.13% BA, and 0.1% BA resulted in mean percentage germination significantly higher than the 75 mM NaCl by 25.3% (in all 3), not significantly different from that of the 0 mM NaCl. In conclusion, concentration of cell-free supernatant and NaCl level influence the effect of Bacillus amyloliquefaciens strain EB2003A cell-free supernatant on mean percentage germination and mean radicle length of corn and soybean.
Article
Full-text available
GDSL esterases/lipases (GELPs), present throughout all living organisms, have been a very attractive research subject in plant science due mainly to constantly emerging properties and functions in plant growth and development under both normal and stressful conditions. This review summarizes the advances in research on plant GELPs in several model plants and crops, including Arabidopsis, rice, maize and tomato, while focusing on the roles of GELPs in regulating plant development and plant–environment interactions. In addition, the possible regulatory network and mechanisms of GELPs have been discussed.
Article
The study investigated the response characteristics of algal-bacterial granular sludge (ABGS) under salinity stress (0 % → 2 %). At 1 % salinity, the sludge performance was inhibited, while recovered rapidly, indicating the ABGS exhibited resistance. However, at 2 % salinity, the suppressed performances did not recover until the stress was eliminated. Under salinity stress, the nutrient removal capacity of the system and the composition and chemical characteristics of extracellular polymers substances also changed. Meanwhile, the ABGS formed adaptation to salinity stress in the early coping process. As a result, the effect of the second 2 % salinity on ABGS was significantly weakened. High-throughput sequencing results showed that the microbial community in ABGS shifted under salinity stress, and the halophilic bacteria genera Arcobacter, Denitromonas, Azoarcus, etc. were enriched, which might be the genetic basis of the adaptation.