ArticlePDF Available

Isolation and characterization of glyphosate-degrading bacteria from different soils of Algeria

Authors:

Abstract and Figures

Glyphosate (N-phosphonomethylglycine) is the most commonly used herbicide worldwide. Due to the concern regarding its toxicity for non-targeted species in soil, finding glyphosate-degrading microorganisms in soil is of interest. The success of this will depend on isolating bacteria with the ability to grow in presence of glyphosate. Five bacterial strains were isolated from different untreated soils of Algeria, the strains were able to grow in a medium containing glyphosate as sole carbon or phosphorus source by enrichment cultures of these soils. Based on 16S rRNA gene sequence analysis, MALDI-TOF MS and biochemical properties, the best strain amongst them (Arph1) was identified as Pseudomonas putida. This isolate showed the highest growth level in the presence of glyphosate as sole phosphorus source. Arph1 was therefore used for further studies for optimization of cultivation conditions for an efficient glyphosate use. The best result of growth was on 1 g/L of glyphosate in minimal medium supplemented with glutamate with initial pH 9.0 at 30°C at 150 rpm within 168 h. Microbial growth during the study was monitored by measuring the optical density at 620 nm. Arph1 was able to tolerate up to 9 g/L of glyphosate. These results show that the bacterial strain may possess potential to be used in bioremediation of glyphosate-contaminated environments.
Content may be subject to copyright.
Vol. 7(49), pp. 5587-5595, 11 December, 2013
DOI: 10.5897/AJMR2013.6080
ISSN 1996-0808 ©2013 Academic Journals
http://www.academicjournals.org/AJMR
African Journal of Microbiology Research
Full Length Research Paper
Isolation and characterization of glyphosate-degrading
bacteria from different soils of Algeria
Ouided Benslama and Abderrahmane Boulahrouf*
Laboratoire de Génie Microbiologique et Applications, Campus Chaâberssas, Faculté des Sciences de la Nature et de la
Vie, Université Constantine 1, Constantine, Algérie.
Accepted 18 November, 2013
Glyphosate (N-phosphonomethylglycine) is the most commonly used herbicide worldwide. Due to the
concern regarding its toxicity for non-targeted species in soil, finding glyphosate-degrading
microorganisms in soil is of interest. The success of this will depend on isolating bacteria with the
ability to grow in presence of glyphosate. Five bacterial strains were isolated from different untreated
soils of Algeria, the strains were able to grow in a medium containing glyphosate as sole carbon or
phosphorus source by enrichment cultures of these soils. Based on 16S rRNA gene sequence analysis,
MALDI-TOF MS and biochemical properties, the best strain amongst them (Arph1) was identified
as Pseudomonas putida. This isolate showed the highest growth level in the presence of glyphosate as
sole phosphorus source. Arph1 was therefore used for further studies for optimization of cultivation
conditions for an efficient glyphosate use. The best result of growth was on 1 g/L of glyphosate in
minimal medium supplemented with glutamate with initial pH 9.0 at 30°C at 150 rpm within 168 h.
Microbial growth during the study was monitored by measuring the optical density at 620 nm. Arph1
was able to tolerate up to 9 g/L of glyphosate. These results show that the bacterial strain may possess
potential to be used in bioremediation of glyphosate-contaminated environments.
Key words: Soil pollution, glyphosate-degrading bacteria, Pseudomonas putida, optimization, cultivation
conditions.
INTRODUCTION
The application of xenobiotic compounds generates
environmental concern by the potential of the unwanted
side effects, as large amounts of substances are
released into the environment. The degradation of xeno-
biotic compounds is an important indicator for healthy
ecosystems. Soil microorganisms can carry out pesticide
degradation and can use the xenobiotic as a source of
carbon, energy and other nutrients to promote microbial
growth (Durkin, 2003). The herbicide glyphosate is often
used to control weeds in grasslands. Despite its
extensive use in Algeria, detailed informations on glypho-
sate degrading-microorganisms are lacking.
Glyphosate (N-phosphonomethylglycine) is the most
commonly used herbicide worldwide (Franz et al., 1997);
it is a broad-spectrum, post-emergence, non-selective
herbicide, that inhibits the enzyme 5-enolpyruvylshikimic
acid-3-phosphate synthase (EPSPS), blocking the syn-
thesis of essential aromatic amino acids (Duke et al.,
2003). The importance of glyphosate degrading-bacteria
has been magnified by the biotechnology application.
Effectively, the bacterial genes encoding for glyphosate-
resistant EPSP synthase were cloned, endowed with
*Corresponding author. E-mail: boulahroufabderrahmane@yahoo.fr. Tel: (00 213) 555 136 100.
5588 Afr. J. Microbiol. Res.
chloroplast transit signals and used to transform plants
(Della-Cioppa et al., 1987) to enable them to survive
treatment following application of glyphosate. Use of this
herbicide in glyphosate-resistant crops has given farmers
cost-effective and broad-spectrum weed control options.
Several bacterial strains were isolated that were able to
degrade glyphosate; most of these bacteria were isolated
from sites already treated by the herbicide. However,
there are few reports of the isolation of bacteria from
untreated sites and no report of glyphosate degrading-
bacteria isolated from Saharian soils. Previous reports
were mainly focused on the screening of bacteria for their
ability to degrade glyphosate. However, the comprehensive
studies of the physiological regulation in bacterial cells
are rather few (Shushkova et al., 2012). Thus, the opti-
mization of cultivation conditions is important to appre-
ciate this physiological regulation, and the identification of
these conditions will make it possible to know which
factors can be applied for bacteria in soil during biore-
mediation.
The aim of the present work was to isolate and
characterize glyphosate-degrading bacteria by using
enrichment cultures for three different untreated Algerian
soils and the assessment of growth response of the
isolates, as well as the optimization of some abiotic para-
meters for the cultivation of isolated strains, providing
maximal effectiveness of the glyphosate degradation.
MATERIAL AND METHODS
Chemicals and media
The isopropylamine salt of glyphosate known as Roundup®
(containing 450 g active ingredient/L of glyphosate, Monsanto) was
purchased from a local store supplier of agricultural products in
Constantine, Algeria.
For the isolation of bacteria using glyphosate as sole source of
carbon and energy, mineral salt medium 1 (MSM1) was used. The
composition of the medium in gram per liter of distilled water, pH
(7.0 to 7.2) was: KH2PO4 (1.5), Na2HPO4 (0.6), NaCl (0.5), NH4SO4
(2), MgSO4 7H2O (0.2), CaCl2 (0.01) and FeSO4 7H2O (0.001).
Whereas, mineral salt medium 2 (MSM2) was used for the isolation
of bacteria using glyphosate as sole phosphorus source, its
composition in gram per liter of distilled water, pH (7.0 to 7.2) is:
Tris buffer (12), glucose (10), NaCl (0.5), KCl (0.5) NH4SO4 (2),
MgSO4 7H2O (0.2) CaCl2 (0.01) and FeSO4 7H2O (0.001). Both
media were supplemented with filter-st erilized (0.2 μm f ilter)
glyphosate and were used to enrich and isolate glyphosate-
degrading strains.
Experimental soil
Soil specimens were collected in April 2012 from three different
untreated soils. The first sample was an agricultural soil, taken from
the Institute of Field Crops in Constantine located between 7°35'
longitude and 36°23' latitude in the center of eastern Algeria. The
second sample was taken from the forest of Chaâberssas located
in the University of Constantine, Algeria. The third sample was
taken from a sandy field located in the region of Biskra located bet-
between 34°51'01" north latitude and 5°43'40" east longitude in the
north-eastern of Algeria on the northern edge of the Sahara,
Desert. Samples of about 1 kg were taken from the first 15 cm of
depth, pooled and sieved. Samples were air dried and stored in
sterile plastic bags at 4°C until use.
Enrichment and isolation of glyphosate-degrading strains
About 5.0 g of each soil were added to 95 mL of MSM1 or MSM2
medium in 250 mL flasks with the addition of glyphosate at a final
concentration of 0.5 g/L and incubated in the dark at 30°C under
shaking condition (150 rpm) for seven days. A 5 mL volume of
these suspensions were then transferred to fresh MSM1 or MSM2
containing 1g/L glyphosate and incubated for seven days. Three
additional successive transfers were made into media successively
containing 3, 6 and 12 g/L of glyphosate. The appropriate dilutions
of enriched samples were plated on plate count agar supplemented
with 1 g/L glyphosate. The plates were incubated at 30°C for 24 h.
Colonies were picked and purified. The strains Arph1, Arph2, Frglu,
Frph and Bisglu were isolated.
Conventional ident ification
For morphological and physiological studies, Arph1, Arph2, Frglu,
Frph and Bisglu isolates were grown in aerobically atmosphere at
30°C on Columbia Agar 5% sheep-blood media (Biomerieux, la
Balm-les-grottes, France). Apart from morphology, mobility,
catalase, oxidase and Gram reaction, physiological studies were
performed by using API 20E.
MALDI-TOF MS identification
The MALDI-TOF mass spectrometry protein analysis was carried-
out as previously described (Seng et al., 2009). Briefly, a pipette tip
was used to pick one isolated bacterial colony from a culture Agar
plate, and to spread it as a thin film on a MTP 384 MALDI-TOF
target plate (Bruker Daltonics, Leipzig, Germany). Twelve (12)
distinct deposits were done for each isolates from 12 different
colonies. Each smear was overlaid with 2 μL of matrix solution
(saturated solution of alpha-cyano-4-hydroxycinnamic acid) in 50%
acetonitrile, and 2.5% tri-fluoracetic-acid, and allowed to dry for five
minutes. Measurements were performed with a Microflex
spectrometer (Bruker).
Spectra were recorded in the positive linear mode for the mass
range of 2 to 20 kDa (parameter settings: ion source 1 (IS1), 20 kV;
IS2, 18.5 kV; lens, 7 kV). A spectrum was obtained after 675 shots
at a variable laser power. The time of acquisition was between 30
sand 1 minper spot. The 12 spectra of the different isolated strains
were imported into the MALDI BioTyper software (version 2.0,
Bruker) and analyzed by standard pattern matching (with default
parameter settings) against the main spectra of 6.213 bacteria in
the BioTyper database. For every spectrum, 100 peaks at most
were taken into account and compared with spectra in the
database.
16S rRNA gene amplification and sequencing
The 16S rRNA gene of the isolates was amplified using the primer
pair fD1-P2 (Weisburg et al., 1991). PCR amplifications were
carried-out in a 50 µL volume containing 5 µL template, 50 mM KCl,
1.5 mM MgCl2, 200 µM each dNTP, 0.2 µM each oligonucleotide
primers and 0.5 units of Taq DNApolymerase (EuroblueTaq,
Eurobio, Les Ulis, France). The thermal cycle consisted of an initial
5 min denaturation at 95°C followed by 35 cycles of 30 s
denaturation at 95°C, primer hybridization at 52°C for 30 s and
elongation at 72°C for 1 min and a final 5-min extension step at
72°C. PCR reactions were examined by electrophoresing 5 µL of
PCR product on a 1% agarose gel stained with ethidium bromide.
The gel was visualized using Gel Doc 1000 (Bio-Rad, California,
USA). Successful PCRs were transferred into PCR purification plate
(Macherey Nagel HOERDT, France) filtrated with vacuum manifolds
Millipore and agitated with a plate (Heidolph instrument Titramax
100). Purified PCR products were sequenced with the use of a
BigDye® Terminator v1.1 Cycle Sequencing Ready Reaction Kit
(Applied Biosystems), the Bvd1, 5X Sequencing Buffer and the
primers 536F, 536R, 800F, 800R, 1050F and 1050R. Sequencing
was then performed in ABI3700 automated capillary sequencer
(Applied Biosystems, Foster City, California, United States).
The nucleotide sequences were edited using ChromasPro 1.34
software (Copyright (c) 2003-2006 by Technelysium Pty Ltd).
Phylogenetic relationships of the genes were reconstructed using
neighbor-joining implemented in MEGA 5 software (Tamura et al.,
2011).
Glyphosate utilization patterns of the different isolates
Inoculums were prepared for each isolate by growing the strains in
50 mL of nutrient broth for three days at 30°C under shaking
condition (150 rpm) till the growth reached late exponential phase.
Cells were harvested by centrifugation at 4, 600 g for 5 min,
washed with 0.9% sterile saline and were re-suspended to a 0.5
McFarland nephelometer standard (Optical density of 0.18 at 625
nm) and this was then used as the inoculum.
Growth experiments with glyphosate as the sole source of carbon
or phosphorus were performed in 250 mL Erlenmeyer flasks
containing 100 mL sterile MSM 1 or MSM 2 with 1 g/L of
glyphosate. A 2 mL of each isolate was inoculated and triplicate
cultures were incubated on a rotary shaker at 150 rpm for 168 h at
30°C. Non-inoculated media served as control. Samples (2 mL)
were withdrawn periodically from the cultures to determine growth
by measurement of the turbidity at 625 nm using a spectro-
photometer.
Optimization of cultivation conditions
To optimize growth in glyphosate enriched media, some important
abiotic factors were chosen. The important factors and their
optimized ranges that were chosen in this experiment were
nutriments (yeast extract, glutamate and glycerol), temperature (30,
37 and 40°C), medium pH (5.0, 6.0, 7.0, 8.0, 9.0 and 10.0) and
initial concentration of glyphosate (1, 3, 5, 7, 9, 12 and 15 g/L). All
experiments were performed in 250 mL flasks containing 100 mL of
MSM2 supplemented with an appropriate amount of glyphosate,
adjusted to an appropriate initial pH and inoculated with 2 mL of
Arph1 strain. The flasks were then incubated at the appropriate
temperature in the dark for seven days and stirred on a rotary
shaker at 150 rpm. Controls without inoculation were kept in similar
conditions. Bacterial growth was followed by taking a sample of 2
mL of cultures after every 24 h until 168 h of incubation and the
optical density was measured at 625 nm.
In the first step, the flasks containing the minimum media were
supplemented with 0.1%, w/v of various nutrients (yeast extract,
glutamate and glycerol). In the second step, minimum medium
supplemented with 0.1%, w/v of glutamate were incubated at
different temperatures (30, 37 and 40°C). In the third step, minimum
Benslama and Boulahrouf 5589
medium supplemented with 0.1%, w/v of glutamate were adjusted
to different initial pH (5.0, 6.0, 7.0, 8.0, 9.0 and 10.0) and incubated
at 30°C. Finally, minimum medium supplemented with 0.1%, (w/v)
of glutamate with different concentrations of glyphosate (1, 3, 5, 7,
9, 12 and 15 g/L), adjusted to pH 9 were incubated at 30°C.
RESULTS
In the present study, five bacterial strains were isolated
from different untreated soils of Algeria; these isolates
have shown an ability to grow in a culture medium in the
presence of the herbicide glyphosate as sole source of
carbon or phosphorus. These isolates were named as
follows: Arph1 and Arph2 isolated from the agricultural
soil of Constantine in the medium containing glyphosate
as sole phosphorus source; Frglu isolated from the forest
soil of Constantine in the medium containing glyphosate
as sole C source; Frph isolated from the forest soil of
Constantine in the medium containing glyphosate as sole
P source and Bisglu isolated from the Saharan soil of
Biskra in the medium containing glyphosate as sole C
source.
Strains identification
For the MALDI-TOF analysis, the obtained score of the
isolate Arph1, was 2.5 close to the species
Pseudomonas putida. The sequence of 1500 bp of the
gene 16S rRNA was deposited in GenBank under
Accession number KC582298. The phylogenetic tree
showing the result of the 16S rRNA of Arph1 is
represented in Figure 1. The sequence of the 16S rRNA
gene of the isolate was 99.5% similar to the 16S rRNA
gene of P. putida (GenBank Accession No. gb
|EU439424.1|) and 99.4% similar to 16S rRNA gene of P.
putida (GenBank Accession No. gb |EU439425.1|). The
result of the analysis of 16S rRNA is consistent with that
of MALDI-TOF, morphological and biochemical properties
(Table 1). Therefore, the isolate was identified as P.
putida.
MALDI-TOF scores obtained for the isolates Arph2 and
Frglu were 2.4 and 2.3 near to Enterobacter cloacae,
respectively. For the analysis of 16S rRNA gene, the
isolate Arph2 showed a sequence similarity of 99.1% with
E. cloacae (GenBank Accession No. gb |JF772064.1|)
and 98.9% with E. cloacae (GenBank Accession No. gb
|CP003737.1|) and Pantoea agglomerans (GenBank
Accession No. gb |AY335552.1|) while the isolate Frglu
showed a similarity of 99.3% with E. cloacae (GenBank
Accession No. gb |JX307682.1|) and 98.8% with E.
cloacae (GenBank Accession No. gb |EF059833.1|). The
sequences of 1513 and 1288 bp of the gene 16S rRNA of
the isolates Arph2 and Frglu were deposited in GenBank
under the Accession numbers KC582299 and KC582300,
respectively. Phylogenetic analysis of 16S rRNA gene
5590 Afr. J. Microbiol. Res.
Figure 1. Neighbor joining tree based on 16S rRNA sequences for Enterobacteriacae. To examine the confidence
of NJ tree, 1000 bootstrap replicates were used.
clustered the two isolates with E. cloacae species as
shown in Figure 1. Based on the results of phylogenetic
and phenotypic tests, the isolates Arph2 and Frglu were
identified as Enterobacter cloacae.
The result of MALDI-TOF identified the isolate Arph3 as
Rahnella aquatilis with a high score of 2.5. The result of
the 16S rRNA gene analysis showed that the isolate Frph
shares a similarity of 99.2% with R. aquatilis (GenBank
Accession No. gb |FJ405361.1|) and 98.8% with R.
aquatilis (GenBank Accession No. emb |X79937.1|). The
sequence of 1337 bp of the gene 16S rRNA of the isolate
was deposited in GenBank under Accession number
KC582301. The result of the analysis of 16S rRNA gene
of the Frph strain is consistent with that of MALDI-TOF,
morphological and biochemical properties (Table 1).
Therefore, the isolate was identified as R. aquatilis.
MALDI-TOF identified the isolate Bisglu as Serratia
marcescens with a high score of 2.4. The result of the
16S rRNA gene analysis showed that the isolate Bisglu
shares a similarity of 99.3% with S. marcescens
(GenBank Accession No. gb |JQ308606.1|) and 98.9%
with S. marcescens (GenBank Accession No. dbj
|AB594756.1|). The sequence of 1233 bp of the gene
16S rRNA of the isolate was deposited in GenBank under
Accession number KC582302. Based on the results of
phylogenetic and phenotypic tests, the isolate Bisglu can
be identified as S. marcescens.
Glyphosate utilization patterns of the different
isolates
As shown in the Figure 2a, of the three isolates grown in
the media containing the glyphosate as sole phosphorus
source, and tested for the growth by measuring their
turbidimetry at 625 nm, P. pudida showed the highest
growth level (OD average = 0.129) suggesting extensive
use of glyphosate. This was followed by E. cloacae (OD
average = 0.100) and R. aquatilis (OD average = 0.084).
However, for the isolates grown in the media containing
the glyphosate as sole carbon source, the growth was
very low as showed in the Figure 2b. The ODs averages
Benslama and Boulahrouf 5591
Table 1. Morphological and biochemical properties of the different isolates. (+), strains positive; (-),
strains negative
Property
Arph2
Frglu
Bisglu
Morphology
Rod-
shaped
Rod-
shaped
Rod-
shaped
Motility
+
+
+
Biochemical tests
Gram test
-
-
-
Oxidase/ catalase
-/+
-/+
-/+
β-galactosidase
+
+
+
Voges-Proskauer
+
+
+
Nitrate production
+
+
+
Lysine decarboxylase
+
+
+
Ornithine decarboxylase
+
+
+
H2S production
-
-
-
Urease
-
-
-
Tryptophan deaminase
-
-
-
Indole production
-
-
-
Gelatinase
-
-
+
Arginine dihydrolase
-
-
-
Citrate
+
+
+
Sugar use
Mannose
+
+
+
+
Glucose
+
+
+
+
Sorbitol
+
+
+
+
Rhamnose
+
+
+
-
Sucrose
+
+
+
+
Melibiose
+
+
+
+
Amygdalin
+
+
+
+
Arabinose
+
+
+
-
Inositol
-
-
-
+
were about 0.051 and 0.045 for S. marscecens and E.
cloacae, respectively.
Effect of abiotic factors
Figure 3 shows the effect of certain nutrients (yeast
extract, glycerol and glutamate) on bacterial growth. As
seen in Figure 3 in the presence of glutamate, growth
kinetic of P. putida strain shows a steady increase in
growth after 24 h of incubation and reached a maximum
of growth of 0.250 after 168 h of incubation, which
represent an increase of over 15% compared to its
growth in the medium without any nutriment. While in the
presence of glycerol and yeast extract, the growth
reached a maximum of 0.245 and 0.206 after 168 h of
incubation, respectively. Therefore, the glutamate was
selected as a carbon source for further studies.
Figure 4 shows the evolution of bacterial growth at
different temperatures (30, 37 and 40°C). A significant
increase in the growth of P. putida strain was noted at
30°C, where the growth of the strain reached its peak of
0.248 after 168 h of incubation. A less significant growth
was obtained at 37°C, reached a maximum of 0.145 after
168 h of incubation. At 40°C, the strain showed a slower
and weak growth and reached its maximum growth of
0.07 after 168 h of incubation. Therefore, the temperature
30°C was selected for further studies.
The effect of pH on the growth of P. putida is shown in
Figure 5. In general, the growth of the strain is greater in
alkaline pH ranging from 7 to 10 over 168 h. When the
initial pH is lower than 7, the growth of the strain
gradually decreased with the decrease of pH. At pH 9,
the growth peaked significantly and reached a maximum
of 0.261 within 168 h of incubation.
The growth kinetic at various initial concentrations of
glyphosate is shown in Figure 6. Increase in microbial
growth was observed till initial concentration of gly-
phosate was increased to 3 g/L. As the concentration of
5592 Afr. J. Microbiol. Res.
0
0.05
0.1
0.15
0.2
0.25
0.3
0
24
48
72
96
120
144
168
Absorbance (625 nm)
Time of incubation (hours)
P. putida
E. cloacae
R. aquatilis
Figure 2a. Growth kinetics of P.pudida, E. cloacae, R.aquatilis strains in
glyphosate as sole phosphorus source.
0
0.05
0.1
0.15
0.2
0.25
0.3
0
24
48
72
96
120
144
168
Absorbance (625 nm)
Time of incubation (hours)
S. marscecens
E.cloacae
Figure 2b. Growth kinetics of E. cloacae and S. marscecens strains in
glyphosate as sole carbon source.
0
0.05
0.1
0.15
0.2
0.25
0.3
0
24
48
72
96
120
144
168
Absorbance (625 nm)
Time of incubation (hours)
glutamate
glycerol
yeast extract
Figure 3. Growth kinetics of P. putida strain in glyphosate as sole
phosphorus source with different nutrients.
Benslama and Boulahrouf 5593
0
0.05
0.1
0.15
0.2
0.25
0.3
0
24
48
72
96
120
144
168
Absorbance (625 nm)
Time of incubation (hours)
30°C
37°C
40°C
Figure 4. Growth kinetics of P. putida strain in glyphosate as sole
phosphorus source supplemented with glutamate (0.1% w/v) in
different temperatures.
0
0.05
0.1
0.15
0.2
0.25
0.3
0
24
48
72
96
120
144
168
Absorbance (625 nm)
Time of incubation (hours)
pH 5
pH 6
pH 7
pH 8
pH 9
pH 10
Figure 5. Growth kinetics of P. putida strain in glyphosate as sole
phosphorus source supplemented with glutamate (0.1% w/v) with
different pH at 30°C.
Figure 6. Effect of the initial concentration of glyphosate on the growth
of P. putida strain.
5594 Afr. J. Microbiol. Res.
glyphosate increased, there was a decrease in the
growth of the isolate. The high concentrations of gly-
phosate severely inhibit bacterial growth.
The highest growth was observed at 1 g/L, which is the
least tested concentration of glyphosate. After 24 h of
incubation, the growth of P. putida in the medium
containing 1 g/L of glyphosate increased significantly and
reached a maximum of 0.265 after 186 h of incubation.
However, No inhibition of growth was observed when
initial gly-phosate concentration was increased further,
indicating that the isolate can tolerate up to 9 g/L of
glyphosate.
DISCUSSION
The limited number of strains isolated from the medium
containing glyphosate as sole carbon or phosphorus
source, is in agreement with the reports of Quinn et al.
(1988), Santos and Flores (1995) and Kryzsko-Lupicka
and Orlik (1997), which showed a significant reduction in
microbial population when glyphosate was added to the
medium culture. This result can be explained by the
toxicity of artificial media due to the mode of action of
glyphosate (the way of shikimic acid is ubiquitous in
microorganisms (Glyphosate makes the organism unable
to synthesize essential aromatic amino acids). In
addition, Liu et al. (1991), Dick and Quinn (1995) report
that when glyphosate is supplied as carbon source,
microbial growth is rare, but growth stimulation is more
apparent when applied in high concentrations. It was
found that the commonly isolated glyphosate-degraders
in the laboratory are the Pseudomonas spp. bacteria
(Jacob et al., 1988; Dick and Quinn, 1995). Five
pseudomonas species isolated that grew solely on
glyphosate were identified (P. maltophilia, P. putida, P.
aeruginosa and Pseudomonas sp.) that whose growth
were not inhibited due to a glyphosate-resistant EPSPS
(Schulz et al., 1985). As well, Bacillus megaterium (Quinn
et al., 1989), Alcaligenes sp. (Tolbot et al., 1984)
Flavobacterium sp. (Balthazor and Hallas, 1986),
Geobacillus caldoxylosilyticus (Obojska et al., 2002),
Rhizobium sp. and Agrobacterium sp. (Liu et al., 1991),
R. aquatilis (Peng et al., 2012) and Enterobacter cloacae
(Kryuchkova et al., 2013) have been reported as
degrading glyphosate. However, there is no report of the
isolation of Serratia marcescens as glyphosate degrading
bacteria. Thus, this finding adds to the list of glyphosate-
degrading bacteria a new degrading species that can be
used in further studies.
The use of microorganisms for bioremediation requires
an understanding of all physiological, microbiological,
ecological, biochemical and molecular aspects involved
in pollutant transformation (Iranzo et al., 2001). The effect
of abiotic factors on bacterial growth was used to
optimize the cultivation conditions that affect significantly
the glyphosate degradation and the bacterial growth. P.
putida was selected in this study because it showed the
highest growth potential in comparison with the other
species as showed in Figure 2a and b. The growth
kinetics of P. putida was monitored over time at 620 nm
using the MSM2 enriched with glyphosate as the sole
source of phosphorus, varying abiotic conditions of the
environment.
Growth rate was most important in the medium
supplemented with glutamate; this results is in agreement
with the report of Shushkova et al. (2012) who shows an
effective glyphosate degradation in the presence of
glutamate in the medium. Kumar and Philip (2006)
reported that the addition of auxiliary carbon to the
system having xenobiotic compounds increased the
biodegradation potential of bacterial culture which was
often because of increase in metabolic activity of the
microbes involved. As reported by Mallick et al. (1999)
and Guha et al. (1997), the co-metabolism appears to
occur commonly in nature. Microbial activity increases
with increasing temperature up to an optimum value. This
result can be attributed to the fact that, at low tem-
perature, the growth of the strain P. putida and the
reaction catalyzed by the enzyme degrading glyphosate
are increased. This indicates that the strain is psy-
chrophilic nature (Patel et al., 2012). Moorman (1994)
stated that within the range of temperature conditions
normally encountered in cultivated soils, the rate of
pesticide degradation generally increased with tem-
perature. Walker et al. (1992) considered soil tem-
perature to be the most important environmental factor
influencing pesticide degradation rate in soils. It has been
reported, that the incubation temperature does not only
affect the pesticide degradation rate but also affect the
growth of the strain.
Slightly alkaline pH is favorable for glyphosate
degradation by the strain, probably due to the increased
bioavailability and the decreased toxicity of glyphosate,
and to the optimal metabolic activity of the bacterial cells.
Singh et al. (2003a, b) report that in soils with higher pH a
higher copy numbers of organo-phosphate degrading
(opd) gene are found, suggesting that the activity of the
enzymes degrading organo-phosphate compounds is
more important at alkaline pH. Decrease in cell density at
high concentrations of glyphosate can be attributed to the
toxicity and the stress of glyphosate on strain. This can
also be explained by the fact that at high concentrations,
the appropriate catabolic enzymes may be repressed.
Another plausible explanation is that the strain may need
an acclimation period to induce the necessary degra-
dative path (Tang and You, 2011). A similar resultwas
found by Moneke et al. (2010) when testing different
initial concentrations of glyphosate on Acetobacter sp. and
and P. fluorescens. Tolerance to high pesticide concen-
trations is critical, since concentrations at contaminated
sites may be several orders of magnitude higher than the
recommended usage doses for these products.
Conclusion
This study is the first report of isolation and charac-
terization of a soil-borne bacterial strains (P. putida, E.
colacae, R. aquatilis and S. marcescens) from an
agricultural, Saharan and forest soil in Algeria that
possess the capacity to use glyphosate. The capacity of
these isolates to survive and grow in the presence of high
concentrations of the herbicide, show that these strains
may possess potential to be used in bioremediation of
glyphosate-contaminated environments or moreover, can
contribute on creating glyphosate-resistant crops.
In addition, this work adds to the list of glyphosate-
degrading bacteria a new degrading species that is S.
marcescens. This study provides also important infor-
mation on optimization of critical parameters of cultivation
conditions to enhance glyphosate degradation by P.
Putida strain.
ACKNOWLEDGEMENT
Authors gratefully acknowledge Professor Michel
Drancourt from Aix Marseille University, Unité des
Rickettsies, Faculté de Médecine, Marseille, France.
REFERENCES
Balthazor TM, Hallas LE (1986). Glyphosate degrading microorganisms
from industrial activated sludge. Appl. Environ. Microbiol. 51:432-434.
Della-Cioppa G, Bauer SC, Taylor ML, Rochester DE, Klein BK, Shah
DM, Fraley RT, Kishore GM (1987). Targeting a herbicide-resistant
enzyme from Escherichia coli to chloroplasts of higher plants. Bio.
Technol. 5:579-584.
Dick RE, Quinn JP (1995). Glyphosate-degrading isolates from
environmental samples: occurrence and pathways of degradation.
Appl. Environ. Microbiol. 43:545-550.
Duke SO, Baerson SR, Rimando AM (2003). Herbicides: glyphosate. In:
Plimmer JR, Gammon DW, Ragsdale NN (Eds.), Encyclopedia of
Agro-chemicals, John Wiley and Sons, New York.
Durkin P (2003). Glyphosate - human health and ecological Risk
assessment. Final Report. USDA, Forest Service.
Franz JE, Mao MK, Sikorski JA (1997). Glyphosate: a unique global
herbicide. In: American Chemical Society Monograph 189, American
Chemical Society, Washington, DC.
Guha A, Kumari B, Roy MK (1997). Possible involvement of plasmid in
degradation of malathion and chlorpyrifos by Micrococcus sp. Folia
Microbiol. Praha. 42:574-576.
Iranzo M, Sain-Pardo I, Boluda R, Sanchez J, Mormeneo S (2001). The
use of microorganisms in environmental remediation. Ann. Microbiol.
51:135-143.
Jacob GS, Gabrow JR, Hallas LE, Kimack NM, Kishore GM, Schaefer J
(1988). Metabolism of Glyphosate in Pseudomonas sp. Strain LBr.
Appl. Environ. Microbiol. 54:2953-2958.
Kryuchkova YV, Burygin GL, Gogoleva NE, Gogolev YV, Chernyshova
MP, Makarov OE, Fedorov EE, Turkovskaya OV (2013). Isolation and
characterization of a glyphosate-degrading rhizosphere strain,
Enterobacter cloacae K7. Microbiol. Res. S0944-5013(13)00034-7.
Kryzsko-Lupicka T, Orlik A (1997). The use of glyphosate as the sole
source of phosphorus or carbon for the selection of soil-borne fungal
strains capable to degrade this herbicide. Chemosphere 34:2601-
2605.
Benslama and Boulahrouf 5595
Kumar M, Philip L (2006). Enrichment and isolation of a mixed bacterial
culture for complete mineralization of endosulfan. J. Environ. Sci.
Health B. 41:81-96.
Liu CM, McLean PA, Sookdeo CC, Cannon FC (1991). Degradation of
the herbicide glyphosate by members of the family Rhizobiaceae.
Arch. Microbiol. 57:1799-1804.
Mallick K, Bharati K, Banerji A, Shakil NA, Sethunathan N (1999).
Bacterial degradation of chlorpyrifos in pure culture and in soil. Bull.
Environ. Contam. Toxicol. 62:48-55.
Moneke AN, Okpala GN, Anyanwu CU (2010). Biodegradation of
glyphosate herbicide in vitro using bacterial isolates from four rice
fields. Afr. J. Biotechnol. 9:4067-4074.
Moorman TB (1994). Pesticide degradation by soil microorganisms:
environmental, ecological and management effects. Adv. Soil. Sci.
33:21-165.
Obojska A, Ternana NG, Lejczak B, Kafarski P, McMullan P (2002).
Organophosphate utilization by the thermophile Geobacillus
caldoxylosilyticus T20. Appl. Environ. Microbiol. 68:2081-2084.
Patel V, Cheturvedula S, Madamwar D (2012). Phenanthrene
degradation by Pseudoxanthomonas sp. DMVP2 isolated from
hydrocarbon contaminated sediment of Amlakhadi canal, Gujarat,
India. J. Hazard. Mater. 30:43-51.
Peng RH, Tian YS, Xiong AS, Zhao W, Fu XY, Han HJ, Chen C, Jin XF,
Yao QH (2012). A novel 5-enolpyruvylshikimate-3-phosphate
synthase from Rahnella aquatilis with significantly reduced
glyphosate sensitivity. Biochem. 35: 4199-210.
Quinn JP, Peden JMM, Dick RE (1988). Glyphosate tolerance and
utilization by the microflora of soils treated with the herbicide. Appl.
Microbiol. Biotechnol. 29:511-516.
Quinn JP, Peden JMM, Dick RE (1989). Carbon-phosphorus bond
cleavage by gram-positive and gram-negative soil bacteria. Appl.
Microbiol. Biotechnol. 31:283-287.
Santos A, Flores M (1995). Effects of Glyphosate on Nitrogen Fixation
of Free-Living Heterotrophic Bacteria. Lett. Appl. Microbiol. 20:349-
352.
Schulz A, Krüper A, Amrhein N (1985). Differential sensitivity of
bacterial 5-enolpyruvylshikimate-3-phosphate synthases to the
herbicide glyphosate. FEMS Microbiol. Lett. 28:297-301.
Seng P, Drancourt M, Gouriet F, La SB, Fournier PE, Rolain JM, Raoult
D (2009). Ongoing revolution in bacteriology: routine identification of
bacteria by matrix-assisted laser desorption ionization time-of-flight
mass spectrometry. Clin. Infect. Dis. 49:543-551.
Shushkova TV, Ermakova IT, Sviridov AV, Leontievsky AA (2012).
Biodegradation of glyphosate by soil bacteria: Optimization of
cultivation and the method for active biomass storage. Microbiol.
81:44-50.
Singh BK, Walker A, Grayston J (2003a). Degradation of chlorpyrifos
and its effects on the soil biota. Pesticide in Air, Plant, Soil and Water
System. In: Del Re AAM, Capri E, Padovani L, Trevisan M (Eds),
Proceeding of the XII Symposium Pesticide Chemistry, Piacenza,
Italy.
Singh BK, Walker A, Morgan JAW, Wright DJ (2003b). Effect of soil pH
on the biodegradation of chlorpyrifos and isolation of a chlorpyrifos-
degrading bacterium. Appl. Environ. Microbiol. 69:5198-5206.
Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S (2011).
MEGA5: Molecular Evolutionary Genetics Analysis using Maximum
Likelihood, Evolutionary Distance, and Maximum Parsimony
Methods. Mol. Biol. Evol. 28:2731-2739.
Tang M, You M (2011). Isolation, identification and characterization of a
novel triazophos-degrading Bacillus sp. (TAP-1). Micr. Res. 167:299-
305.
Tolbot HW, Johnson LM, Munneck DM (1984). Glyphosate utilization by
Pseudomonas sp. and Alcaligenes sp. isolated from environmental
sources. Curr. Microbiol. 10:255-259.
Walker A, Moon YH, Welch SJ (1992). Influence of temperature, soil
moisture and soil characteristics on the persistence of alachlor.
Pestic. Sci. 35:109-116.
Weisburg WG, Barns SM, Pelletier DA, Lane DJ (1991). 16S ribosomal
DNA amplification for phytopathogenic study. J. Bacteriol. 173:697-
703.
... In soils, glyphosate is primarily converted to AMPA by the action of the enzyme glyphosate oxidoreductase (GOX), while the enzyme C-P lyase, produces sarcosine and phosphate (Feng et al., 2020;Zhan et al., 2018). A number of bacteria capable of metabolising glyphosate through the action of either of these enzymes have been isolated from contaminated soil samples, including members of the genera Achromobacter, Agrobacterium, Alcaligenes, Aminobacter, Arthrobacter, Bacillus, Comamonas, Enterobacter, Flavobacterium, Geobacillus, Lysinibacillus, Ochrobactrum, Pseudomonas, Sinorhizobium and Streptomyces (Firdous et al., 2020;Gorodylova et al., 2021;Hove-Jensen et al., 2014;Ouided & Abderrahmane, 2013;Singh et al., 2020;Villegas et al., 2018). Members of Arthrobacter, Bacillus, Sinorhizobium and Pseudomonas genera have also been associated with AMPA degradation (Firdous et al., 2020;Hove-Jensen et al., 2014;Ouided & Abderrahmane, 2013;Singh et al., 2020;Zhan et al., 2018). ...
... A number of bacteria capable of metabolising glyphosate through the action of either of these enzymes have been isolated from contaminated soil samples, including members of the genera Achromobacter, Agrobacterium, Alcaligenes, Aminobacter, Arthrobacter, Bacillus, Comamonas, Enterobacter, Flavobacterium, Geobacillus, Lysinibacillus, Ochrobactrum, Pseudomonas, Sinorhizobium and Streptomyces (Firdous et al., 2020;Gorodylova et al., 2021;Hove-Jensen et al., 2014;Ouided & Abderrahmane, 2013;Singh et al., 2020;Villegas et al., 2018). Members of Arthrobacter, Bacillus, Sinorhizobium and Pseudomonas genera have also been associated with AMPA degradation (Firdous et al., 2020;Hove-Jensen et al., 2014;Ouided & Abderrahmane, 2013;Singh et al., 2020;Zhan et al., 2018). ...
Article
Full-text available
The herbicide glyphosate has several potential entry points into composting sites and its impact on composting processes has not yet been evaluated. To assess its impact on bacterial diversity and abundance as well as on community composition and dynamics, we conducted a mesocosm experiment at the Montreal Botanical Garden. Glyphosate had no effect on physi-cochemical property evolution during composting, while it was completely dissipated by the end of the experiment. Sampling at Days 0, 2, 28 and 112 of the process followed by 16S rRNA amplicon sequencing also found no effect of glyphosate on species richness and community composition. Differential abundance analyses revealed an increase of a few taxa in the presence of glyphosate, namely TRA3-20 (order Polyangiales), Pedo-sphaeraceae and BIrii41 (order Burkholderiales) after 28 days. In addition, five amplicon sequence variants (ASVs) had lower relative abundance in the glyphosate treatment compared to the control on Day 2, namely Coma-monadaceae, Pseudomonas sp., Streptomyces sp., Thermoclostridium sp. and Actinomadura keratinilytica, while two ASVs were less abundant on Day 112, namely Pedomicrobium sp. and Pseudorhodoplanes sp. Most differences in abundance were measured between the different sampling points within each treatment. These results present glyphosate as a poor determinant of species recruitment during composting.
... The ability of these microbes to be viable in higher and even lower pH (i.e below 5.0) depends on their capability to maintain the difference in pH of their intracellular and extracellular environment [13]. Bacillus cereus CB4 [14] showed optimum glyphosate degradation at pH 6.0 -7.0, while a more alkaline condition was preferred by Pseudomonas putida, with optimum growth at pH 9.0 [15]. However, inadequate literature is available on glyphosatedegrading bacteria favoring acidic conditions since most glyphosate degraders prefer neutral-alkaline pH for optimum degradation rate [6]. ...
... Ochobactrum sp. GDOS [16] and Pseudomonas putida [15] have the highest bacterial growth at 30 °C, suggesting extensive utilization of glyphosate by these bacteria. The reason for the high preferred temperature by Pseudomonas sp. in this study ( Fig. 4.5a) could be a result of the high temporal region in which this bacterium was isolated. ...
Article
Full-text available
Bioremediation of pollutants, such as herbicides, is an economic and environmentally friendly process. Glyphosate is an active ingredient in most herbicides utilized for weed control and desiccation on cereal and other grain crops globally. Glyphosate pollution poses a threat to the environment and the habitats in it. In this study, an isolated molybdenum-reducing bacterium was characterized for its potential to degrade glyphosate and utilized as the sole source of carbon and electron donor. The effects of incubation time, glyphosate concentration (carbon source), inoculum size, pH, temperature, aeration and heavy metals on the growth of this bacterium were spectrophotometrically assayed as OD600 nm. The bacterium degrades glyphosate faster under shaking conditions, optimally at pH 7.0, concentration 1.0 g/L, temperature 40 ºC, and inoculum size 400 µL. Growth of this bacterium was significantly inhibited by heavy metals in the order of Cu>Zn>Pb>Hg>Ag>Fe compared to the control. Glyphosate can serve as an electron donor source in hexavalent molybdenum reduction, but poorly supports molybdenum blue (Mo-blue) production compared to glucose. The dual role of this isolate as a metal reducer and glyphosate degrader makes it unique already and an important instrument for the bioremediation of mixed pollutants.
... Ochrobactrum sp. GDOS (Hadi et al., 2013) and P. putida (Ouided and Abderrahmane, 2013) showed maximum bacterial growth at 30˚C, recommending vast utilization of GLX by these bacteria. For the meantime, B. cereus CB4 revealed optimum biodegradation at 35˚C (Fan et al., 2012b). ...
Article
Full-text available
Glyphosate (GLX) is a crucial organophosphonate herbicide applied to eliminate unwanted, grasses and herbaceous plants in various vegetation management situations. Its widespread and consequently application is causing environmental pollution, there is a need to remove it from the environment using an eco-friendly and cost-effective method. Biodegradation is an effective method for the removal of GLX. But various environmental factor effecting is degradation after its applications such as aerobic and anaerobic environment, different pH and temperature, soil water contents and soil textures. Glyphosate resistant bacterial strains have been isolated from GLX-contaminated soil which were proficient to utilize GLX as a carbon source. High Pressure Liquid Chromatography instrument was used to determine the rate of GLX biodegradation in liquid media and soil under different environmental conditions. The five out of eleven strains such as WAG2 (Serratia liquefaciens), WAG4 (Klebsiella variicola), WAG45 (Enterobacter cloacae), WAG9 (Pseudomonas aeruginosa), and WAG11 (Enterobacter ludwigii) were proficient in average degraded 95 to 98 mg kg-1 glyphosate in soil followed by liquid media 93 to 96 mg kg-1 within 28 days of incubation in different environmental factors. The optimal circumstances for the degradation of GLX were found to be aerobic, pH 7, 40˚C, soil water content at 10% and sandy loam texture showed maximum degradation at 100 mg kg-1 compared to 200 mg kg-1 concentration of GLX. This study demonstrates the potential of isolated bacterial strains for efficient degradation of GLX, which can be exploited for remediation of GLX in contaminated soil.
... The use of herbicides in agriculture in Ouargla, a region in the Algerian Sahara, has led to the accumulation of pollutants in the soil. This poses a significant challenge for soil remediation in these arid land ecosystems (Sviridova et al., 2015;Benslama and Boulahrouf, 2013). The biological approach to soil remediation is considered more ecologically friendly and less expensive than physical-chemical techniques (Zhan et al., 2018). ...
Article
Full-text available
Glyphosate, also known as N-phosphonomethylglycine, is the herbicide that is widely used across the globe. As there are concerns over its potential toxicity to non-target soil species, there is a growing interest in identifying glyphosate-degrading microorganisms in soil. Biodegradation, by actinobacteria, is a very promising approach to eliminate this pesticide from contaminated environments. The present work isolated and identified actinobacteria capable of degrading glyphosate from Saharan agriculture, as well as determined how the application of this herbicide affects the abundance of actinobacteria present in soil. It was observed that the use of glyphosate led to an increased abundance of actinobacteria in the soil compared to the untreated soil. Among this population, an actinobacterial strain was isolated from glyphosate contaminated soil by the enrichment method, and was identified to possess the greatest capability to degrade glyphosate at 50 mg/L. The identification of this strain was achieved through a combination of cultural, morphological, biochemical, and molecular techniques. This included the use of 16S rDNA sequencing, leading to its successful classification as Streptomyces sp. strain SRH22. This strain was assigned the accession number OQ302556 by the National Center for Biotechnology Information (NCBI). A rapid, sensitive, and straightforward spectrophotometric technique was employed for the quantification of glyphosate. Results showed that the optimal biodegradation (90.2%) was obtained under a temperature of 30 degrees, a PH of 7.2, and an inoculum volume of 4% timed over six days. This work shows that the Streptomyces SRH22 presents good potentiality to be used as a bioremediation agent for agricultural soils in the Algerian Sahara.
... The herbicide acted as a sole source of carbon and energy for isolated microorganisms, and its concentration was chosen based on literature data (Benslama and Boulahrouf, 2013;Fan et al., 2012;McAuliffe et al., 1990;Nourouzi et al., 2011). Mineral medium consisted of 7.0 g/L Na 2 HPO 4 × 2H 2 O, 2.8 g/L KH 2 PO 4 , 0.5 g/L NaCl, 1.0 g/L NH 4 Cl. ...
Article
The applicability of herbicidal ionic liquids (HILs) as an alternative form of herbicides is currently evaluated. Yet, the available research is lacking information on the behaviour of herbicidal ionic liquids upon addition to the environment, i.e., if cations and anions act as separate moieties or remain an ionic salt. Hence, we tested degradation of five HILs with the glyphosate anion, their bioavailability in soil, toxicity towards microorganisms, impact on the biodiversity and the abundance of phnJ and soxA genes. The cations were proven to be slightly or moderately toxic. The properties of cations determined the properties of the whole formulation, which it might suggest that cations and anion act as the independent mixture of ions. The mineralisation efficiencies were in the range of 15–53%; however, in the case of cations (except non-toxic choline), only 13–20% were bioavailable for degradation. The hydrophobic cations were proven to be highly sorbed, while the anion was readily available for microbial degradation regardless of its counterion. The approach to enrich test samples with isolated microorganisms specialised in glyphosate degradation resulted in higher degradation efficiencies, yet not high enough to mitigate the negative impact of cations. In addition, increased activity of enzymes participating in glyphosate degradation was observed. In the view of obtained results, the use of cationic surfactants in HILs structure is not recommended, as sorption was shown to be determining factor in HILs degradation efficiency. Moreover, obtained results indicate that corresponding ions in HILs might act as separate moieties in the environment.
... Fresh bacterial inoculum was prepared by cultivating in TSB tubes to an optical density of 600 nm and adding 1 mL of inoculum equivalent to (10 7 CFU/mL) from the six bacterial strains in 250 mL of MS broth containing 50 mg/L of CAP and FBD mixture according to the method of Gao et al. [20] under different growth conditions (temperatures: 25, 30, 35, and 40 °C; pH degrees: 6.0, 7.0, 8.0, and 9.0; NaCl concentrations: 0, 0.5, 2.5, and 5.0%; 0, control means that no salt is added to the MS broth during its preparation; and interval incubation periods: 3, 7, 11, and 16 days at the agitation of 150 rpm using a shaking incubator). The optical density (OD) at 600 nm (OD 600) was determined by using a spectrophotometer (UV-2101/3101 PC; Shimadzu Corporation, Analytical Instruments Division, Kyoto, Japan) following the modifications described by Ouided and Abderrahmane [36] and John et al. [37]. Referring to the one variable at a time (OVAT) technique, all factors were held constant in these experiments, with only the targeted variable being changed for the optimization of process variables [22,38]. ...
Article
Full-text available
This study aimed to isolate, purify, and identify some bacteria from different sources known to be contaminated with pesticides and evaluate their ability to degrade two important pesticides, chlorantraniliprole (CAP), and flubendiamide (FBD). In our study, six isolates showed maximum growth in the presence of CAP and FBD in the growth media as a sole carbon source. The isolates were purified and then identified by biochemical and morphological tests, MALD-TOF-MS, and 16S rRNA techniques, as Bacillus subtilis subsp. subtilis AZFS3, Bacillus pumilus AZFS5, Bacillus mojavensis AZFS15, Bacillus paramycoides AZFS18, Pseudomonas aeruginosa KZFS4, and Alcaligenes aquatilis KZFS11. The degradation ability of studied bacterial strains against pesticides was estimated under different conditions (temperatures, pH, salt, and incubation time). The results reveal that the optimal conditions for all bacterial strains’ growth were 30–35 °C, pH 7.0, 0.0–0.5% NaCl, and an incubation period of 11 days at 150 rpm in the presence of diamide insecticides at 50 mg/L. The capacity of six bacterial strains of CO2 production and degradation ability against various diamide pesticides and other pesticide groups (Profenofos, Cypermethrin, Carbofuran, and Malathion) were evaluated. The results show that the Pseudomonas aeruginosa KZFS4 (LC599404.1) strain produced the highest CO2 content, about 1.226 mg CO2/16 day, with efficacy in the biodegradation of FBD-CAP (78.6%), while the absorbance of bacterial growth (OD 600) on various pesticides ranged from 1.542 to 1.701. Additionally, Consortium-(No. 3)-mix-6-strains gave 1.553 mg CO2/16 days with efficacy (99.6%) and turbidity of 2.122 to 2.365 (OD 600) on various pesticides. In conclusion, the six bacterial strains could play an important role in the biodegradation process of pollutants in soils.
Article
Laccase is a multicopper enzyme that plays a unique role in bioremediation of environmental pollutants. Bacteria were isolated from hospital wastewater and screened for laccase production. The laccase production process condition was optimised, and the laccase obtained was characterised. The 16S rRNA molecular analysis conducted on the best laccase producer revealed a Bacillus sp. NU2 identified. The process conditions: pH5, 45 oC, 100 rpm, 5% inoculum, and growth constituents viz: tangerine peel and wheat bran agro-wastes, beef extract, ammonium persulfate, glucose, galactose, xylose, sorbitol, fructose carbon sources; and 4-aminophenol inducer optimally stimulated laccase production. The Bacillus sp. NU2 laccase was optimal at pH and temperature conditions of 8.0 and 60 oC, with a noteworthy pH and thermal stability observed. Furthermore, NU2 laccase showed a moderate/high tolerance and relative activity effect on various chemical inhibitors, halides and surfactant of triton x-100 (105 ± 0.92%), PMSF (107 ± 0.81%), and NaCl (94 ± 0.81%) at 1, 3, and 6 (mM) concentration. Additionally, NU2 laccase maintained a relative activity of 101%, 104%, and 102% for Mg2+, Zn2+, and Fe3+ at 1, 3, and 6 mM respectively. Acetone and propanol significantly upregulated laccase activity at 114 ± 0.0008% and 118.24 ± 0.35 and also at 30 and 20 (%) concentrations. Conclusively, the tolerant effect of Bacillus sp. NU2 laccase in pH, temperature, inhibitors and organic solvents suggests its potential for biotechnological application and promotion of a greener environment.
Article
Full-text available
Enterobacter species are considered to be an opportunistic human pathogen owing to the existence of antibiotic-resistant strains and drug resides; however, the detailed analysis of the antibiotic resistance and virulence features in environmental isolates is poorly characterized. Here, in the study, we characterized the biochemical characteristics, and genome, pan-genome, and comparative genome analyses of an environmental isolate Enterobacter sp. S-16. The strain was identified as Enterobacter spp. by using 16S rRNA gene sequencing. To unravel genomic features, whole genome of Enterobacter sp. S-16 was sequenced using a hybrid assembly approach and genome assembly was performed using the Unicycler tool. The assembled genome contained the single conting size 5.3 Mbp, GC content 55.43%, and 4500 protein-coding genes. The genome analysis revealed the various gene clusters associated with virulence, antibiotic resistance, type VI secretion system (T6SS), and many stress tolerant genes, which may provide important insight for adapting to changing environment conditions. Moreover, different metabolic pathways were identified that potentially contribute to environmental survival. Various hydrolytic enzymes and motility functions equipped the strain S-16 as an active colonizer. The genome analysis confirms the presence of carbohydrate-active enzymes (CAZymes), and non-enzymatic carbohydrate-binding modules (CBMs) involved in the hydrolysis of complex carbohydrate polymers. Moreover, the pan-genome analysis provides detailed information about the core genes and shared genes with the closest related Enterobacter species. The present study is the first report showing the presence of YdhE/NorM in Enterobacter spp. Thus, the elucidation of genome sequencing will increase our understanding of the pathogenic nature of environmental isolate, supporting the One Health Concept.
Article
Full-text available
The water quality and phytoplankton of Ufiobodo and Ebonyi River reservoirs were studied for nine months from March to November, 2019 to evaluate the suitability of the reservoirs for domestic water supply and habitat for aquatic organisms. Temperature, pH, conductivity, total dissolved solid (TDS), transparency, dissolved oxygen (DO) and depth were measured in situ, while water samples were collected for determination of nitrate, iron, silicate and phosphate in the laboratory. Phytoplankton samples were also collected concurrently using plankton net of mesh size 45μm and mouth diameter (0.26m) and identified in the laboratory using standard keys. The results showed that mean conductivity (32.33μS/cm), TDS (16.00mg/l), nitrate (33.99mg/l), depth (10.36m) and phosphate (4.81mg/l) were higher at Ebonyi than Ufiobodo Reservoir but DO was lower at Ebonyi Reservoir (2.77mg/l) than at Ufiobodo Reservoir (6.67mg/l). Mean DO level was below the permissible limit for drinking water at Ebonyi Reservoir while nitrate, phosphate and iron concentrations exceeded permissible limits in both Reservoirs. Five phytoplankton Phyla were encountered, predominated by Chlorophyta (107 individuals/l, H=2.80 and d=3.85) and Bacillariophyta (81 individuals/l, H=2.39 and d=2.50) at Ufiobodo and Ebonyi Reservoirs, respectively. Principal Component Analysis (PCA) showed that temperature, conductivity, TDS, depth, DO, nitrate and phosphate were the major factors influencing phytoplankton abundance and diversity of the reservoirs. Thus, the reservoirs should be protected to maintain/improve the water quality for domestic use and to support the socio-economic and ecological services they provide.
Article
Full-text available
Conditions for obtaining the active biomass of Ochrobactrum anthropi GPK 3 and Achromobacter sp. Kg 16, bacteria which are able to degrade the herbicide glyphosate (N-phosphonomethylglycine), were investigated. In the batch culture, degradation was most effective in the medium with pH 6.0–7.0 and aeration at 10–60% of air saturation supplemented with glutamate and ammonium chloride as sources of carbon and nitrogen, respectively. Due to the adaptation of the cells and induction of the relevant enzymatic systems, the inoculum grown in the presence of glyphosate exhibited 1.5–2-fold higher efficiency of xenobiotic degradation than that grown with other sources of phosphorus (orthophosphate and methylphosphonic acid). The efficiency of the toxicant decomposition increased with an increase in a specific load of glyphosate, which the cells were subjected to during the initial stage of growth. The specific load was regulated both by the initial cell concentration and the concentration of the phosphorus source, and the effect was probably determined by its availability to microorganisms. Storage of the liquid biopreparation as a paste with stabilizers (ascorbate, thiourea, and glutamate) at room temperature for 50 days resulted in high level of bacteria viability and a degrading activity approximately equal to that obtained when the bacteria were maintained on the agar medium containing glyphosate at 4°C with monthly transfers to the fresh culture medium.
Article
Full-text available
Due to their high metabolic diversity and high adaptability, microorganisms are able to live in the most varied of "natural" and "artificial" habitats created by environmen- tal contamination. Different microbes can use a great variety of refractory pollutants, thus permitting their use in ex and in situ bioremediation. The implementation of biotreatment processes requires the use of evaluation methods of pollution and the success of bioreme- diation. The microorganisms can also be used as biosensors. Laboratory microcosm exper- iments would allow isolation of new microorganisms, assessment of the methods for eval- uating pollution and the risk of implementation of non- and recombinant microorganisms. Bioremediation technologies have an open future linked to multidisciplinary scientific work.
Article
Full-text available
Plant-growth-promoting rhizobacteria exert beneficial effects on plants through their capacity for nitrogen fixation, phytohormone production, phosphate solubilization, and improvement of the water and mineral status of plants. We suggested that these bacteria may also have the potential to express degradative activity toward glyphosate, a commonly used organophosphorus herbicide. In this study, 10 strains resistant to a 10mM concentration of glyphosate were isolated from the rhizoplane of various plants. Five of these strains - Alcaligenes sp. K1, Comamonas sp. K4, Azomonas sp. K5, Pseudomonas sp. K3, and Enterobacter cloacae K7 - possessed a number of associative traits, including fixation of atmospheric nitrogen, solubilization of phosphates, and synthesis of the phytohormone indole-3-acetic acid. One strain, E. cloacae K7, could utilize glyphosate as a source of P. Gas-liquid chromatography showed that E. cloacae growth correlated with a decline in herbicide content in the culture medium (40% of the initial 5mM content), with no glyphosate accumulating inside the cells. Thin-layer chromatography analysis of the intermediate metabolites of glyphosate degradation found that E. cloacae K7 had a C-P lyase activity and degraded glyphosate to give sarcosine, which was then oxidized to glycine. In addition, strain K7 colonized the roots of common sunflower (Helianthus annuus L.) and sugar sorghum (Sorghum saccharatum Pers.), promoting the growth and development of sunflower seedlings. Our findings extend current knowledge of glyphosate-degrading rhizosphere bacteria and may be useful for developing a biotechnology for the cleanup and restoration of glyphosate-polluted soils.
Article
Full-text available
The 5-enolpyruvylshikimate-3-phosphate synthase (EPSPS; EC 2.5.1.19) is a key enzyme in the shikimate pathway for the production of aromatic amino acids and chorismate-derived secondary metabolites in plants, fungi, and microorganisms. It is also the target of the broad-spectrum herbicide glyphosate. Natural glyphosate resistance is generally thought to occur within microorganisms in a strong selective pressure condition. Rahnella aquatilis strain GR20, an antagonist against pathogenic agrobacterial strains of grape crown gall, was isolated from the rhizosphere of grape in glyphosate-contaminated vineyards. A novel gene encoding EPSPS was identified from the isolated bacterium by complementation of an Escherichia coli auxotrophic aroA mutant. The EPSPS, named AroA(R. aquatilis), was expressed and purified from E. coli, and key kinetic values were determined. The full-length enzyme exhibited higher tolerance to glyphosate than the E. coli EPSPS (AroA(E. coli)), while retaining high affinity for the substrate phosphoenolpyruvate. Transgenic plants of AroA(R. aquatilis) were also observed to be more resistant to glyphosate at a concentration of 5 mM than that of AroA(E. coli). To probe the sites contributing to increased tolerance to glyphosate, mutant R. aquatilis EPSPS enzymes were produced with the c-strand of subdomain 3 and the f-strand of subdomain 5 (Thr38Lys, Arg40Val, Arg222Gln, Ser224Val, Ile225Val, and Gln226Lys) substituted by the corresponding region of the E. coli EPSPS. The mutant enzyme exhibited greater sensitivity to glyphosate than the wild type R. aquatilis EPSPS with little change of affinity for its first substrate, shikimate-3-phosphate (S3P) and phosphoenolpyruvate (PEP). The effect of the residues on subdomain 5 on glyphosate resistance was more obvious.
Article
The potent inhibition of the shikimate pathway enzyme 5-enolpyruvylshikimate-3-phosphate (EPSP) synthase by the broad-spectrum herbicide glyphosate (N-[phosphonomethyl]glycine) was confirmed for the enzymes extracted from various bacteria, a green alga and higher plants. However, 5 out of 6 species belonging to the genus Pseudomonas were found to have EPSP synthases with a 50- to 100-fold decreased sensitivity to the inhibitor. Correspondingly, growth of these 5 species was not inhibited by 5 mM glyphosate, and the organisms did not excrete shikimate-3-phosphate in the presence of the herbicide.
Article
5 – Enolpyruvylshikimate – 3 – phosphate (EPSP) synthase is a chloroplast–localized enzyme of the shikimate pathway in plants. The enzyme is synthesized as a cytoplasmic precursor with an N–terminal transit peptide sequence that post–translationally directs import into chloroplasts. EPSP synthase is of significant agronomic importance because it is the primary target for glyphosate, the active ingredient in the herbicide. Roundup®. A gene from Escherichia coli encoding a mutant, glyphosate resistant form of EPSP synthase was fused behind a portion of the cDNA that encodes the transit peptide sequence of Petunia hybrida EPSP synthase. A chimeric enzyme, produced by in vitro transcription and translation of the gene fusion, was rapidly imported into chloroplasts and proteolytically processed to give a stable, glyphosate resistant enzyme.
Article
Alachlor degradation in soils followed first-order reaction kinetics. Half-lives in moist soil varied from 11.30 to 34.8 days at 25°C to 95.9 to 279.6 days at 5°C, and at 15°C varied from 15.7 to 83.1 days at 5 kPa soil water stress (field capacity) to 82.8 to 281.4 days at 1500 kPa (permanent wilting point). Degradation rates in laboratory incubations with fluctuating temperatures were predicted with reasonable accuracy from the constant-temperature data. The degradation rate and extent of adsorption were lower in subsoils than in soils from the plough layer. Degradation rate was positively correlated with microbial biomass and microbial respiration, and adsorption was positively correlated with soil organic matter content. Persistence of alachtor in field plots was correlated well with variations in weather pattern during the period September 1990 to July 1991, with an effective half-life varying from 20 to 60 days. Persistence in the field plots was predicted accurately by a computer model of herbicide behaviour.
Article
The effect of the herbicide glyphosate (N-(phosphonomethyl)glycine) on the growth, respiration and nitrogen fixation of Azotobacter chroococcum and A. vinelandii was studied. Azotobacter vinelandii was more sensitive to glyphosate toxicity than A. chroococcum. Recommended dosages of glyphosate did not affect growth rates. More than 4 kg ha-1 is needed to find some inhibitory effect. Specific respiration rates were 19.17 mmol O2 h-1 g-1 dry weight for A. chroococcum and 12.09 mmol h-1 g-1 for A. vinelandii. When 20 kg ha-1 was used with A. vinelandii, respiration rates were inhibited 60%, the similar percentage inhibition A. chroococcum showed at 28 kg ha-1. Nitrogen fixation dropped drastically 80% with 20 kg ha-1 in A. vinelandii and 98% with 28 kg ha-1 in A. chroococcum. Cell size as determined by electron microscopy decreased in the presence of glyphosate, probably because glyphosate induces amino acid depletion and reduces or stops protein synthesis.
Article
The herbicide glyphosate was used as a selection agent for isolation of fungal strains capable to degrade phosphorus-to-carbon bond from standard sandy-clay soil. The studies have shown that the herbicide used in Martin medium as a sole source of phosphorus br carbon caused the decrease of the fungal population and substantially changed strain composition, thus selecting those which are able to degrade glyphosate.