ArticlePDF Available

Biodegradation of naphthalene, BTEX, and aliphatic hydrocarbons by Paraburkholderia aromaticivorans BN5 isolated from petroleum-contaminated soil

Authors:

Abstract and Figures

To isolate bacteria responsible for the biodegradation of naphthalene, BTEX (benzene, toluene, ethylbenzene, and o-, m-, and p-xylene), and aliphatic hydrocarbons in petroleum-contaminated soil, three enrichment cultures were established using soil extract as the medium supplemented with naphthalene, BTEX, or n-hexadecane. Community analyses showed that Paraburkholderia species were predominant in naphthalene and BTEX, but relatively minor in n-hexadecane. Paraburkholderia aromaticivorans BN5 was able to degrade naphthalene and all BTEX compounds, but not n-hexadecane. The genome of strain BN5 harbors genes encoding 29 monooxygenases including two alkane 1-monooxygenases and 54 dioxygenases, indicating that strain BN5 has versatile metabolic capabilities, for diverse organic compounds: the ability of strain BN5 to degrade short chain aliphatic hydrocarbons was verified experimentally. The biodegradation pathways of naphthalene and BTEX compounds were bioinformatically predicted and verified experimentally through the analysis of their metabolic intermediates. Some genomic features including the encoding of the biodegradation genes on a plasmid and the low sequence homologies of biodegradation-related genes suggest that biodegradation potentials of strain BN5 may have been acquired via horizontal gene transfers and/or gene duplication, resulting in enhanced ecological fitness by enabling strain BN5 to degrade all compounds including naphthalene, BTEX, and short aliphatic hydrocarbons in contaminated soil.
This content is subject to copyright. Terms and conditions apply.
1
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
www.nature.com/scientificreports
Biodegradation of naphthalene,
BTEX, and aliphatic hydrocarbons
by Paraburkholderia
aromaticivorans BN5 isolated from
petroleum-contaminated soil
Yunho Lee, Yunhee Lee & Che Ok Jeon
To isolate bacteria responsible for the biodegradation of naphthalene, BTEX (benzene, toluene,
ethylbenzene, and o-, m-, and p-xylene), and aliphatic hydrocarbons in petroleum-contaminated
soil, three enrichment cultures were established using soil extract as the medium supplemented with
naphthalene, BTEX, or n-hexadecane. Community analyses showed that Paraburkholderia species
were predominant in naphthalene and BTEX, but relatively minor in n-hexadecane. Paraburkholderia
aromaticivorans BN5 was able to degrade naphthalene and all BTEX compounds, but not n-
hexadecane. The genome of strain BN5 harbors genes encoding 29 monooxygenases including two
alkane 1-monooxygenases and 54 dioxygenases, indicating that strain BN5 has versatile metabolic
capabilities, for diverse organic compounds: the ability of strain BN5 to degrade short chain aliphatic
hydrocarbons was veried experimentally. The biodegradation pathways of naphthalene and BTEX
compounds were bioinformatically predicted and veried experimentally through the analysis of
their metabolic intermediates. Some genomic features including the encoding of the biodegradation
genes on a plasmid and the low sequence homologies of biodegradation-related genes suggest that
biodegradation potentials of strain BN5 may have been acquired via horizontal gene transfers and/
or gene duplication, resulting in enhanced ecological tness by enabling strain BN5 to degrade all
compounds including naphthalene, BTEX, and short aliphatic hydrocarbons in contaminated soil.
Over the last few decades, diverse organic pollutants including petroleum fuels, pesticides, solvents, pharma-
ceuticals, and other organic chemicals have been produced industrially and released intentionally or uninten-
tionally during their transports or storages1,2. ese organic pollutants are generally highly persistent, have low
degradability, and can become trapped for long periods of time in the soil minerals of contaminated sites, and
subsequently bioaccumulated, due to their hydrophobic and stable chemical properties35. Because they have
potentially adverse eects on human health and multiple environments (aquatic, terrestrial, and atmospheric),
the contamination of these organic pollutants is a cause of great environmental concerns, which necessities the
need to restore contaminated sites1,610.
Bioremediation, a technique that employs microorganisms to remove organic pollutants, has been proven
to be sustainable, eco-friendly, and cost-eective among remediation technologies for contaminated sites1114.
Many diverse microorganisms that have good pollutant-degrading ability have been previously isolated through
enrichment processes using synthetic basal media, which has led to extensive genomic and metabolic studies on
pollutant biodegradation1520. However, bioremediation of contaminated sites by these microorganisms oen
results in failure due to unfavorable several biotic and abiotic factors including low degradability, low viabil-
ity, low pollutant availability, depletion of nutrients, and unfavorable pH, oxygen, temperature and moisture at
contaminated sites2123. erefore, it has been suggested that the use of microorganisms mainly responsible for
pollutant degradation at contaminated sites is requisite for successful bioremediation of contaminated sites15,2430.
Therefore, the identification and characterization of key players responsible for the degradation of organic
Department of Life Science, Chung-Ang University, Seoul, 06974, Republic of Korea. Correspondence and requests
for materials should be addressed to C.O.J. (email: cojeon@cau.ac.kr)
Received: 23 April 2018
Accepted: 16 November 2018
Published: xx xx xxxx
OPEN
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
2
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
pollutants at contaminated sites are of necessity, and enrichment cultures mimicking contaminated environ-
ments are good approaches to identify and isolate microorganisms having an ability to actively degrade organic
pollutants at contaminated sites27,3135.
To date, biodegradation studies have been mostly focused on the isolation and characterization of micro-
organisms having an ability to degrade single organic compounds although environments are generally con-
taminated with mixed organic compounds33,36,37. erefore, a microorganism with a wide range of degradation
abilities on mixed organic pollutants or a consortium of microorganisms with a degradation ability on each
component of mixed organic pollutants is required for the successful bioremediation of contaminated sites. Soil
of gas station is oen contaminated with mixed organic compounds including BTEX (benzene, toluene, ethyl
benzene, and o-, m-, p-xylene), polycyclic aromatic hydrocarbons (PAH), and aliphatic hydrocarbons by leak-
ing gasoline and diesel fuels. ese compounds have been recognized as environmental priority pollutants that
should be removed from polluted environments because they are toxic, genotoxic, mutagenic or carcinogenic to
human6,8. erefore, to remediate gas station soil contaminated with gasoline and diesel fuels, microorganisms
able to degrade BTEX, PAH, and aliphatic hydrocarbons may be necessary. However, bacterial members harbor-
ing degradation potentials for dierent types of compounds such as PAH, BTEX, and aliphatic hydrocarbons
have been very rarely reported. It was reported that Rhodococcus strains showing broad degradation capabilities
toward various compounds, including naphthalene, BTEX, aliphatic hydrocarbons were successfully isolated, but
their genomic and biochemical features for their biodegradation were not investigated38. In this study, to identify
and isolate bacteria mainly responsible for the biodegradation of BTEX, naphthalene, and aliphatic hydrocarbons
from soil contaminated with gasoline and diesel fuels, we established three enrichment cultures using soil extract,
supplemented with naphthalene-, BTEX-, or hexadecane, as an enrichment medium to mimic environmen-
tal conditions and isolated a bacterial strain, Paraburkholderia aromaticivorans BN5, which had naphthalene-,
BTEX-, and short chain aliphatic hydrocarbon-biodegrading ability. In addition, in this study we investigated the
genomic and biochemical features of strain BN5 for the biodegradation of naphthalene-, BTEX-, and short chain
aliphatic hydrocarbons.
Results
Enrichment cultures and microbial community analysis. e use of microorganisms that have good
pollutant-degrading ability in the contaminated site of interest, not just in the laboratory or in other sites, is a
prerequisite for successful bioremediation. In this study, to isolate bacteria mainly responsible for the biodegra-
dation of naphthalene, BTEX, and aliphatic hydrocarbons in a contaminated site, soil extract was prepared using
soil from a gas station that had been contaminated with gasoline and diesel fuels for a long period of time. e
contaminated soil extract, rather than a dened minimal medium, which has been commonly used previously for
the enrichment of pollutant-degrading bacteria, was used as an enrichment medium.
e enrichment cultures supplemented with naphthalene, BTEX mixture, and n-hexadecane as the sole car-
bon source were sub-cultured three times, and the bacterial communities of the contaminated soil used for the
enrichment and the nal enrichment cultures were analyzed by PacBio RS II sequencing. e bacterial commu-
nity analysis showed that members of the genera Paraburkholderia and Rhodococcus were present at low rela-
tive abundances (approximately 0.9 and 0.3%, respectively) in the contaminated soil (Fig.S1). However, aer
the enrichments, the genus Paraburkholderia became predominant in both enrichment cultures supplemented
with naphthalene and BTEX, accounting for 98.4% and 79.3% of their total bacterial communities, respectively
(Fig.1). On the other hand, the enrichment culture supplemented with n-hexadecane was dominated by the
genus Rhodococcus, with approximately 49.8% relative abundance. e genus Paraburkholderia was also identied
from the enrichment culture supplemented with n-hexadecane, but its relative abundance was only 4.0%. ese
data suggest that the enrichments of naphthalene, BTEX, and hexadecane degraders were successfully accom-
plished and the bacterial group belonging to the genus Paraburkholderia may degrade naphthalene, BTEX, and
aliphatic hydrocarbons.
The genera Azotobacter, Ralstonia, Nitrobacter, Rhodanobacter, Thermomonas, Hydrogenophaga, and
Curvibacter were identied from the enrichment cultures supplemented with BTEX or n-hexadecane as minor
bacterial groups present.
Isolation of major bacteria from the enrichment cultures and their biodegradation abili-
ties. To isolate major bacteria with naphthalene-, BTEX-, and hexadecane-degrading abilities from the enrich-
ment cultures, the nal enrichment culture samples were spread on R2A agar and bacterial strains belonging to
the genera Paraburkholderia and Rhodococcus, the dominant bacterial genus groups in the enrichment cultures,
were isolated. Bacterial strains belonging to the genus Paraburkholderia were isolated from all three enrichment
cultures supplemented with naphthalene, BTEX, and n-hexadecane, while bacterial strains belonging to the genus
Rhodococcus were isolated from only the enrichment culture supplemented with n-hexadecane. Bacterial strains
belonging to the genera Ralstonia, Rhodanobacter, and Bacillus that were identied as minor bacterial groups
in the enrichment cultures were also successfully isolated, but the isolation of minor bacterial groups such as
Azotobacter, Nitrobacter, and ermomonas from the enrichment cultures failed.
e naphthalene-, BTEX-, and n-hexadecane-biodegrading abilities of bacterial isolates from the enrichment
cultures were evaluated in sterile 160 ml serum bottles. Interestingly, all Paraburkholderia strains regardless of
their isolation source had the ability to degrade naphthalene and all BTEX compounds, but they did not degrade
n-hexadecane despite the fact that they were also isolated from the enrichment culture supplemented with
n-hexadecane (data not shown). Rhodococcus strains isolated from the enrichment culture supplemented with
n-hexadecane clearly showed n-hexadecane degradation ability, but they did not degrade naphthalene and BTEX
compounds (data not shown). However, all other minor bacterial isolates did not exhibit the ability to degrade
naphthalene, BTEX, or n-hexadecane (data not shown). Although it was reported that some Paraburkholderia
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
3
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
strains may have an biodegradation ability for naphthalene or BTEX compounds, Paraburkholderia strains
showing biodegradation abilities for naphthalene or BTEX compounds and moreover for both naphthalene and
all BTEX compounds have not yet been reported. erefore, in this study, we investigated the metabolic and
genomic features of a Paraburkholderia strain, strain BN5, which showed the ability to biodegrade naphthalene
and all BTEX compounds.
Biodegradation of naphthalene and BTEX by strain BN5 in minimal salt basal media (MSB)
and soil slurry systems. The ability of strain BN5 to degrade naphthalene, BTEX compounds, and
n-hexadecane was further evaluated over time in MSB (Fig.2). e degradation tests clearly showed that strain
BN5 had the ability to degrade naphthalene as well as all six BTEX compounds. However, strain BN5 did not
degrade n-hexadecane, even aer 14 days of incubation (data not shown). Naphthalene degradation occurred
more quickly than BTEX compound degradation by strain BN5, potentially because 30 mg/l of naphthalene was
singly present in the test serum bottles, while 30 mg/l of each six BTEX compounds (total 180 mg/l) were present
together in the test serum bottles.
Figure 1. Bacterial community compositions of the nal enrichment cultures supplemented with naphthalene,
BTEX mixture (benzene, toluene, ethyl benzene, and o-, m-, p-xylene xy1:1:1:1:1:1), and hexadecane. e
bacterial 16S rRNA gene sequences were classied at the genus level using the mothur soware against the
SILVA Gold reference database. “Others” represents taxa that comprised <1% of the total reads in three
samples.
Figure 2. Biodegradation of naphthalene and BTEX (benzene, toluene, ethyl benzene, and o-, m-, p-
xylene xy1:1:1:1:1:1) compounds by strain BN5 in serum bottles containing naphthalene (30 mg/l) or BTEX
mixture (initial concentration of each BTEX compounds, 30 mg/l) in MSB media. Serum bottles without the
inoculation of strain BN5 were used as negative controls; naphthalene and BTEX decreases in the negative
controls were negligible (not shown). e symbols represent averages of triplicate experiments and the error
bars indicate their corresponding standard deviations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
4
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
e naphthalene and BTEX degradation by strain BN5 was also evaluated in soil slurry systems to better
mimic the contaminated soil environment (Fig.3). e degradation tests demonstrated that the degradation of
naphthalene and BTEX compounds in the soil slurry systems did not occur, or occurred at a very low rate, with-
out treatment with strain BN5 or nutrients (1 g/l of NH4Cl and 0.5 g/l of Na3PO4) during 7 days of incubation,
despite the fact that the soil slurry systems were established using unsterilized soil. However, the biodegradation
of naphthalene and BTEX compounds was signicantly improved by the addition of nutrients. When strain BN5
was inoculated into the soil slurry systems, the biodegradation of the tested compounds was also improved. In the
soil slurry systems that were treated with both nutrients and strain BN5 together, the biodegradation of naphtha-
lene and BTEX compounds was greatly improved. ese results suggest that strain BN5 was metabolically active
in the degradation of naphthalene and BTEX compounds in the slurry conditions mimicking (physiologically and
ecologically) the contaminated site. However, the results also suggest that there is a lack of nutrients for the bio-
degradation of naphthalene and BTEX compounds by bacteria in the contaminated site, and a supply of nutrients
containing nitrogen and phosphorus may be necessary to improve the biodegradation.
Figure 3. Biodegradation of naphthalene (A), benzene (B), toluene (C), ethyl benzene (D), m-, p-xylene (E),
and o-xylene (F) by strain BN5 in soil slurry systems. Symbols in the gures are as follows: untreated (--),
treated with nutrients (--), treated with strain BN5 (--), and treated with nutrients and strain BN5 (--).
Unsterilized freshwater and soil were used for the setting of the soil slurry systems. e symbols represent
averages of triplicate experiments and the error bars indicate their corresponding standard deviations.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
5
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
General features of the whole genome of strain BN5. To investigate the genomic and metabolic fea-
tures of strain BN5, its whole genome was completely sequenced, and its general features summarized in Table1.
e complete genome of strains BN5 consisted of two circular chromosomes of 4,378.9 and 2,992.2 kb and six
circular plasmids of 665.0, 489.4, 162.8, 152.4, 39.5, and 27.5 kb with an average G wiC content of 62.9% (Table1
and Fig.S2). Most of the genomic features of strain BN5, such as the number of chromosomes, total chromo-
some sizes, total gene numbers, protein coding sequences, and total rRNA and tRNA gene numbers are relatively
similar with those of other Paraburkholderia strains3941. A phylogenetic analysis based on the 16S rRNA gene
sequences indicated that strain BN5 formed a phyletic lineage within the genus Paraburkholderia (Fig.S3) and
was most closely related to P. phytormans PsJNT, known to be a plant colonizing bacterium, with a very high 16S
rRNA gene sequence similarity (99.4%). However, the average nucleotide identity (ANI) and in silico DNA-DNA
hybridization (DDH) values of strains BN5 and PsJNT were 88.5% and 36.5%, respectively, clearly lower than the
95% ANI and 70% DDH cut-o values generally accepted for species delineation42,43, suggesting that strain BN5 is
a new species of the genus Paraburkholderia44. e total genome size of strain BN5 (8.9 Mb) was marginally larger
than that of strain PsJNT (8.2 Mb), and while strain BN5 harbors six plasmids, P. phytormans PsJNT contains only
one plasmid of 121 kb40, indicating that strain BN5 may have more diverse and versatile metabolic capabilities.
CRISPR, clustered regularly interspaced short palindromic repeats, along with cas (CRISPR-associated)
genes, are a bacterial defense mechanism against bacteriophage predation and their presence has been gener-
ally accepted as an evidence for previous bacteriophage infections. Genomic analysis showed that the genome
of strain BN5 contains only one possible CRISPR (Table1), while 23 predicted genomic islands (GIs) and 171
predicted transposase genes, indicating lateral gene transfers, were identied. ese ndings suggests that the
genome of strain BN5 has experienced extensive and complex genetic alterations or exchanges by lateral gene
transfers, rather than from bacteriophage infections during its evolutionary history. Strain BN5 harbors genes
encoding 29 monooxygenases and 54 dioxygenases that may be related to the biodegradation of various organic
compounds including naphthalene, BTEX, and other hydrocarbons; among them, 6 monooxygenase and 14 diox-
ygenase genes are located on plasmids, indicating possible inow of the biodegradation genes. Interestingly, strain
BN5 also harbors chromosomal two alkane 1-monooxygenase (alkB) genes (CJU94_RS03555 on chromosome
1 and CJU94_RS23135 on chromosome 2). e presence of alkB genes suggests that strain BN5 may have the
potential to degrade aliphatic hydrocarbons with dierent chain lengths or under specic conditions, although
the strain did not show hexadecane degradation ability in the above degradation tests. erefore, the ability
of strain BN5 to degrade dierent aliphatic hydrocarbons with dierent chain lengths (heptane, nonane, and
n-hexadecane) was evaluated. e biodegradation tests showed that strain BN5 does not degrade n-hexadecane,
in agreement with our previous biodegradation tests, but does have the ability to degrade short chain aliphatic
hydrocarbons such as heptane (C7) and nonane (C9) (Fig.4). In particular, nonane had been completely degraded
within ve days of incubation.
Genome-based prediction of naphthalene and BTEX degradation pathways and their experi-
mental verication. Previous studies have shown that naphthalene and BTEX compounds are metabolized
through a variety of degradation pathways depending on the microorganism39. e catabolic genes and metabolic
pathways of naphthalene and BTEX compounds in strain BN5 were bioinformatically predicted and experimen-
tally veried through gas chromatography/mass spectrometry (GC/MS) analysis of the metabolic intermediates
of naphthalene and each BTEX compound. e catabolic genes of naphthalene and BTEX compounds in strain
BN5 are located on plasmid 2 of 489.4 kb (pBN2), suggesting that strain BN5 probably obtained its degradation
capabilities of naphthalene and BTEX compounds through plasmid transfer.
Characteristic Chromosome (Chr) or plasmid (pBN)
Type Chr1 Chr2 pBN1 pBN2 pBN3 pBN4 pBN5 pBN6
Size (Kb) 4,388.9 2,992.2 665.0 489.4 162.8 152.4 39.5 27.5
G + C content (%) 63.2 63.7 61.3 60.0 59.7 61.5 59.0 59.7
Total genes 3,965 2,632 619 533 166 145 39 32
Protein coding sequences 3,823 2,535 569 483 143 137 37 29
Pseudogenes 76 81 49 50 23 8 2 2
rRNA (16S, 23S, 5S) operons 3 3 — — — — — —
tRNA genes 54 7 1 — — — — 1
Other RNA genes 4 — — — — — —
Genomic islands 18 5
CRISPRs (questionable)*– (1) — ——————
Predicted transposase genes 38 28 17 52 27 9 — —
Monooxygenase genes 14 9 1 4 1 —
Dioxygenase genes 17 23 4 10 — — — —
Alkane 1-monooxygenase genes 1 1 — — — — — —
GenBank accession numbers NZ_CP022989–96.1
Table 1. General features of the chromosomes and plasmids in P. aromaticivorans strain BN5. *CRISPRs with
fewer than three perfect repeats or nonidentical repeats are considered “questionable.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
6
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
It has been generally reported that one of two catabolic pathways, the catechol pathway or the gentisate
pathway, is generally used by bacteria for the biodegradation of naphthalene45. Bioinformatic analysis of the
genome showed that the naphthalene degradation gene cluster in strain BN5 is located on plasmid pBN2 and
its operon structure is similar to those of Alteromonas naphthalenivorans46, Polaromonas naphthalenivorans47,48,
and Ralstonia sp. strain U249 known as the nag operon, involved in metabolizing naphthalene via the gentisate
biochemical degradation pathway (Fig.5A and TableS1). Figure5A and TableS1 show the physical map of the
naphthalene catabolic genes and their associated functions, respectively. Gene homology and putative functions
clearly show that the naphthalene degradation gene cluster contains naphthalene dioxygenase (nagAaAbAcAd),
salicylate-5-hydroxylase (nagGH), and gentisate 1,2-dioxygenase (nagI) genes as key naphthalene metabolic
enzymes that are regulated by a LysR-type regulatory gene (nagR), suggesting that strain BN5 metabolizes naph-
thalene through salicylate and gentisate to fumarate and pyruvate, as do Ralstonia sp. U249, P. naphthalenivorans
CJ248, and A. naphthalenivorans SN246 (Fig.5B). Because salicylate and gentisate are key metabolic intermediates
in the gentisate pathway for naphthalene biodegradation, the presence of these intermediates was investigated
using a gas chromatography-mass spectroscopy detector (GC-MSD) during naphthalene degradation by strain
BN5. GC/MS analysis showed that salicylate and gentisate were detected from naphthalene-grown cell suspen-
sions (Fig.S4), which veried that strain BN5 metabolizes naphthalene via the gentisate biochemical degradation
pathway.
Bioinformatic analysis showed that gene clusters probably responsible for benzene/toluene/xylene degrada-
tion were also found to be located on plasmid pBN2. e benzene/toluene/xylene catabolic genes were split into
two gene clusters with an approximate distance of 18.7 kb (Fig.6A) and their associated functions are described in
Figure 4. Biodegradation of heptane, nonane, and hexadecane by strain BN5 in serum bottles containing
heptane, nonane or hexadecane (100 mg/l) in MSB media. Serum bottles without the inoculation of strain BN5
were used as negative controls and the concentrations of heptane, nonane, and hexadecane were measured aer
ve days of incubation at 30 °C. e bars represent averages of triplicate experiments and the error bars indicate
their corresponding standard deviations.
Figure 5. Physical map of naphthalene degradation genes located on the plasmid pBN2 of strain BN5 (A) and
a proposed biochemical pathway of naphthalene degradation (B). e putative functions of the naphthalene
biodegradation genes were predicted and are presented in TableS1.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
7
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
TableS2. Previously, two primary benzene biodegradation pathways with the conversions of benzene to phenol or
benzene to cis-dihydrobenzenediol as the result of the rst oxidation have been proposed as being typical aerobic
benzene biochemical pathways50,51. No gene encoding benzene monooxygenase for the conversion of benzene to
phenol was identied from the genome of strain BN5. However, it has been reported that toluene monooxygen-
ase responsible for ring monooxidation of toluene and phenol 2-hydroxylase converting phenol to catechol are
also able to oxidize benzene to phenol50. e metabolic gene cluster analysis showed that strain BN5 harbored all
genes encoding toluene 4-monooxygenase (tmoABCDE) and phenol 2-hydroxylase (dmpKLMNOP) (Fig.6A and
TableS2). In addition, strain BN5 harbored genes coding for phenol 2-hydroxylase converting phenol to catechol
and catechol 2,3-dioxygenase (xylE) converting catechol to 2-hydroxymuconate semialdehyde in the gene cluster.
Based on these results, a potential benzene degradation pathway, which has phenol and catechol as intermediates,
in strain BN5 was proposed (Fig.6B). To conrm the benzene biodegradation pathway, the metabolic inter-
mediates of benzene were investigated using a GC-MSD during benzene biodegradation: phenol and catechol
were detected as metabolic intermediates from benzene-grown suspensions (Fig.S5). e detection of phenol
and catechol as metabolic intermediates and the metabolic gene analysis suggest that strain BN5 metabolizes
benzene through a metabolic pathway with rst benzene monooxidation by toluene monooxygenase or phenol
2-hydroxylase and phenol and catechol as metabolic intermediates.
It has been proposed that aerobic toluene biodegradation is typically initiated by ve dierent oxidations,
the methyl group oxidation, ring monooxidation at position 2, 3, or 4, or ring 2,3-dioxidation of toluene, as
the rst step52. e bioinformatic analysis of the benzene/toluene/xylene catabolic genes shows that strain BN5
harbors genes encoding a complete toluene 4-monooxygenase (tmoABCDE), 4-cresol dehydrogenase (pchCF),
and 4-hydroxybenzaldehyde dehydrogenase (pchA) as putative toluene biodegradation genes (Fig.6A and
TableS2), possibly indicating that toluene degradation in strain BN5 is initiated by ring monooxidation at the
para position of toluene to form 4-hydroxytoluene (p-cresol) by TmoABCDE, followed by the conversion of
4-hydroxytoluene to 4-hydroxybenzaldehyde and 4-hydroxybenzoate by PchCF and PchA subsequently. Based on
Figure 6. Physical maps of the metabolic genes probably responsible for benzene, toluene, and xylene
degradation located on the plasmid pBN2 of strain BN5 (A). e metabolic genes are split into two gene
clusters with a distance of approximately 18.7 kb and their putative functions were predicted and presented in
Supplementary TableS2. Based on the predicted functions of the metabolic genes and the conrmation through
GC/MS analysis of the metabolites, the biochemical pathways of benzene (B), toluene (C), and xylene (D)
degradation in strain BN5 were proposed. Genes encoding the enzymes with asterisks were not identied in
the genome of strain BN5 by bioinformatics analysis. e putative functions of the naphthalene biodegradation
genes were predicted and presented in TableS2.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
8
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
the gene annotations, a potential toluene biodegradation pathway of strain BN5 was proposed (Fig.6C). However,
the TmoABCDE- and PchCF-encoding genes were present in dierent gene clusters, indicating that they may be
separately and independently regulated. To conrm the proposed toluene biodegradation pathway, the toluene
biodegradation intermediates were investigated using a GC-MSD: 4-hydroxybenzoate was detected as a meta-
bolic intermediate from toluene-grown cell suspensions (Fig.S6). Because 4-hydroxybenzoate is not a metabolic
intermediate in four other toluene biodegradation pathways, it was concluded that strain BN5 metabolizes tolu-
ene through the toluene 4-monooxygenase pathway.
Figure2 showed that strain BN5 has the ability to degrade all three xylene compounds, which sug-
gests that strain BN5 harbors all xylene degradation genes. However, genes related to the xylene biodegrada-
tion, except for genes encoding benzoate 1,2-dioxygenase (XylXY) converting methyl benzoate (toluate) to
1,2-dihydroxy-methyl-cyclohexa-3,5-diene-carboxylate were not identied in the genome of strain BN5 by bio-
informatic analysis. Generally, it has been reported that biodegradation of xylenes including three structural
isomers of dimethyl benzene (o-, m-, and p-xylenes) is initiated through the oxidation of a methyl substituent
of xylene to 2-methylbenzyl alcohol by xylene monooxygenase or the direct oxidation of the aromatic ring by
xylene dioxygenase5355. In the xylene monooxygenase pathway, 2-methylbenzyl alcohol is subsequently con-
verted to methyl benzaldehyde, methyl benzoate, and 2-dihydroxy-methyl-cyclohexa-3,5-diene-carboxylate by
benzylalcohol dehydrogenase, benzylaldehyde dehydrogenase, and benzoate 1,2-dioxygenase (XylXY), respec-
tively. erefore, based on the presence of XylXY-encoding genes (Fig.6A and TableS2), a possible xylene biodeg-
radation pathway in strain BN5 was proposed (Fig.6D). To conrm the proposed xylene biodegradation pathway,
xylene biodegradation intermediates from xylene-grown cell suspensions using each o-, m-, and p-xylene as a
sole carbon source were investigated through GC/MS analysis. e results showed that methylbenzyl alcohol
and 1,2-dihydroxy-methyl-cyclohexa-3,5-diene-carboxylate were detected as metabolic intermediates from
xylene-grown cell suspensions (Fig.S7), which suggests that strain BN5 may metabolize xylene through the oxi-
dation of a methyl substituent of xylene to 2-methylbenzyl alcohol by xylene monooxygenase. However, the
genome annotation of strain BN5 failed to uncover the genes encoding xylene monooxygenase, benzylalcohol
dehydrogenase, and benzylaldehyde dehydrogenase, which might be caused by their very low sequence homolo-
gies with known corresponding genes.
Two aerobic ethylbenzene metabolic pathways using either the aromatic ring oxidation of ethylbenzene
by ethylbenzene dioxygenase to cis-1,2-dihydroxy-2,3-dihydroethylbenzene56 or the ethyl group oxidation of
ethylbenzene by naphthalene dioxygenase to 1-phenethyl alcohol15,57 have been proposed. Bioinformatic analysis
showed that ethylbenzene dioxygenase or genes homologous with ethylbenzene dioxygenase were not identied
from in the genome of strain BN5 although it carries 29 monooxygenases and 54 dioxygenases, suggesting that
strain BN5 may metabolize ethylbenzene through the ethyl group oxidation of ethylbenzene by naphthalene
dioxygenase. In the ethyl group degradation pathway that uses naphthalene dioxygenase, 1-phenethyl alcohol is
typically converted into 2-hydroxy acetophenone by naphthalene dioxygenase or into benzoylacetate by acetophe-
none carboxylase57,58. erefore, metabolic intermediates of ethyl benzene biodegradation were investigated using
a GC-MSD: during ethylbenzene degradation, 2-hydroxy-acetophenone was detected from ethylbenzene-grown
cell suspensions (Fig.S8). Moreover, a gene encoding acetophenone carboxylase was not identied from the
genome analysis of strain BN5. Based on the analysis of the genome and intermediates of strain BN5, we propose
a possible ethyl benzene biodegradation pathway in strain BN5 that naphthalene dioxygenase catalyzes in multi-
ple steps, which was in common with those in previous reports15,57 (Fig.S9).
Discussion
Bioremediation using bacterial metabolic processes is considered to be one of the most ecient and eco-friendly
clean-up techniques of contaminated environments1114. e microbial catabolic capability for organic pollutants
is a prerequisite for bioremediation practices and thus the isolation and characterization of microorganisms that
have the ability to metabolize a wide range of organic pollutants has been extensively studied11,45,53,5961. To date,
articial synthetic basal media containing a sole organic compound have mostly been used to enrich or isolate
microorganisms that have organic substance-degradation ability on25,26,29,31.
However, the use of articial synthetic basal media as an enrichment medium has oen resulted in the enrich-
ment of microorganisms that only grow successfully in laboratory conditions, not in contaminated sites32,62. In
addition, although the isolated microorganisms have the ability to degrade organic pollutants in laboratory con-
ditions, they oen show a low degradation ability in eld conditions due to various unfavorable factors21,23,63.
erefore, the use of microorganisms that can actively degrade organic substances at contaminated sites is nec-
essary for the successful clean-up of contaminated sites. To identify and isolate microorganisms that eectively
degrade organic pollutants at contaminated sites, various approaches such as enrichment culture33, metagenom-
ics64, stable isotope probing27,65 and uorescence in situ hybridization (FISH)66 have been used. Among them, an
enrichment culture mimicking contaminated environments is one of the easiest and most promising approaches
to identify and isolate microorganisms that have pollutant degradation ability at contaminated sites33,35,6769.
Therefore, to isolate microorganisms responsible for the degradation of BTEX, naphthalene, and aliphatic
hydrocarbons in petroleum (gasoline and diesel)-contaminated soil, soil extract, not an articial synthetic basal
medium, was used as an enrichment medium.
ree enrichment cultures for the isolation of naphthalene-, BTEX-, and aliphatic hydrocarbon-degrading
bacteria from gasoline and diesel fuel-contaminated soil were established. A bacterial group belonging to
Paraburkholderia was present, with 0.9% relative abundance in the contaminated soil, and predominantly
enriched in the enrichment cultures using soil extract supplemented with naphthalene and BTEX (Figs1 and S1).
e degradation tests showed that Paraburkholderia strain BN5 that was isolated from the enrichment cultures
had the ability to degrade naphthalene as well as all BTEX compounds in a slurry system designed to mimic
the contaminated soil conditions (Fig.3), again reinforcing the idea that Paraburkholderia may be responsible
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
9
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
for the biodegradation of naphthalene and BTEX compounds in the contaminated soil. However, many other
biotic and abiotic factors aecting biodegradation activity of Paraburkholderia that we are not able to mimic in
the slurry system are possibly present in contaminated soils21,70. In addition, it is not thought that the growth of
Paraburkholderia with 0.9% of relative abundance in the contaminated soil entirely relies on only the biodegra-
dation of naphthalene and BTEX compounds because contaminated soils contain many other organic carbon
sources. erefore, in situ biodegradation test may be necessary to assess more clearly biodegradation activity of
Paraburkholderia in the contaminated soil.
Paraburkholderia species were also identied from the enrichment culture supplemented with n-hexadecane,
although their relative abundance was low (Fig.1) and Paraburkholderia strains isolated from the enrichment
cultures had no degradation activity for n-hexadecane. However, bioinformatic analysis of the genome of
Paraburkholderia strain BN5 showed that the strain harbors genes encoding two alkane 1-monooxygenases, a key
enzyme in the aliphatic hydrocarbon biodegradation (Table1). e biodegradation test of strain BN5 for aliphatic
hydrocarbons with dierent chain lengths showed that strain BN5 had the ability to biodegrade short chain ali-
phatic hydrocarbons such as heptane and nonane (Fig.4), which might be the reason that Paraburkholderia was
only found in minor abundance in the hexadecane enrichment culture.
Many members belonging to the genera Pseudomonas, Rhodococcus, Mycobacterium, Sphingomonas, and
Burkholderia have been considered to probably possess a broad range of catabolic potentials toward diverse aro-
matic compounds30,39,7173. In particular, it has been largely assumed through the genomic analysis that many
members of Alcaligenaceae, Burkholderiaceae, and Comamonadaceae families have biodegradation potentials
towards a vast array of aromatic compounds, including several priority pollutants39. Auret and colleagues
reported that they successfully isolated Rhodococcus strains showing broad degradation capabilities toward vari-
ous compounds, including naphthalene, BTEX, aliphatic hydrocarbons, and more recalcitrant compounds such
as methyl tert-butyl ether and ethyl tert-butyl ether, but the genomic and biochemical features of the strains for
the biodegradation of the organic compounds were not investigated38. is study is the rst report for the enrich-
ment, isolation, and genomic and biochemical analysis of Paraburkholderia with biodegradation abilities for all of
naphthalene, BTEX, and short chain aliphatic hydrocarbons although there is a report that P. phytormans PsJN,
a close relative of strain BN5, may have a potential degrading-activity for organic compounds30,40,71.
e degradation tests clearly showed that strain BN5 had a biodegradation ability for all of naphthalene,
BTEX compounds, and short chain aliphatic hydrocarbons, but almost biochemical degradation gene clusters,
except for the naphthalene degradation gene cluster, are not well organized (Figs5 and 6 and TablesS1 and S2).
In addition, most of these degradation genes were encoded on a plasmid (plasmid pBN2). ese analyses suggest
that strain BN5 may have recently acquired its biodegradation abilities via horizontal gene transfer (HGT) or the
gain of a plasmid with degradation genes, resulting in successful adaptation to gasoline and diesel-contaminated
soil. Phylogenetic analysis indicated that toluene monooxygenase and alkane 1-monooxygenase genes were not
homologous with those of the genera Paraburkholderia and Burkholderia, indicating that they might have been
introduced via HGT (Fig. S10). Although strain BN5 was able to biodegrade naphthalene, all BTEX compounds,
and short chain aliphatic hydrocarbons, some key degradation genes such as xylene degradation genes were not
identied in the genome of strain BN5.
In this study, we isolated Paraburkholderia strain BN5, capable of degrading naphthalene, BTEX, and short
chain aliphatic hydrocarbons from gasoline and diesel fuel-contaminated soil. e genomic analysis of strain
BN5 suggests that strain BN5 may have acquired its biodegradation capabilities via plasmid gain, HGT, and/or
gene changes under the contaminated soil conditions, resulting in enhanced ecological tness by enabling strain
BN5 to degrade all of naphthalene, BTEX, and short chain aliphatic hydrocarbons. We predicted the degradation
pathways of naphthalene and all BTEX compounds in strain BN5 by analyzing the naphthalene and BTEX degra-
dation gene clusters, and the identied biochemical degradation pathways were experimentally veried through
the analysis of their metabolic intermediates. e availability of strain BN5 and its complete genomic information
will provide us with insights into better understanding of the physiological and genomic features of strain BN5
and its naphthalene-, BTEX-, and aliphatic hydrocarbon-degrading ability in soil environments contaminated
with gasoline and diesel fuels. However, further studies on gene regulation of the biodegradation pathways in
strain BN5 are necessary to understand the simultaneous biodegradation features of naphthalene-, BTEX-, and
aliphatic hydrocarbons by strain BN5 in contaminated soil more clearly.
Methods
Soil samples and enrichment cultures. Soil extract was used as an enrichment medium to isolate
bacteria responsible for degrading pollutants in a contaminated site. For the preparation of soil extract and
the enrichment of naphthalene-, BTEX (benzene, toluene, ethyl benzene, and o-, m-, p-xylene)-, and aliphatic
hydrocarbon-degrading bacteria, soil samples were obtained from a gasoline and diesel fuel-contaminated site
(37°5021.5N 126°5937.2E) of a gas station located in Yangju, Gyeonggi Province, South Korea. Only freshly
collected soil was used for the preparation of the soil extracts, which were prepared according to the procedure
described previously67. In brief, 800 g of the contaminated soil was sieved through a 2 mm mesh and then mixed
with 2 liters of distilled water. e mixture was vigorously agitated for 2 h at room temperature (approximately
25 °C) and centrifuged (5,000 rpm, 25 °C, 10 min). e supernatant was ltered through a 0.45 µm membrane
lter (Millipore, USA) and then supplemented with NH4Cl (1 g/l) and Na3PO4 (0.5 g/l) for nutrients. ree cot-
ton-plugged 500 ml Erlenmeyer asks containing 100 ml of the soil extract and 10 g of the contaminated soil
were prepared, and 0.5 g of naphthalene pellets, 1.0 ml of BTEX mixture (benzene, toluene, ethyl benzene, and
o-, m-, p-xylene yl1:1:1:1:1:1), and 1.0 ml of n-hexadecane were directly added into each of three asks. e three
enrichment cultures were incubated at 25 °C with shaking (180 rpm) and subcultured (1:20) into fresh soil extract
containing the same amount of each compound three times every two weeks.
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
10
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
Bacterial community analysis of enriched cultures. Genomic DNA was extracted from the contam-
inated soil and the nal enrichment cultures using a FastDNA SPIN Kit for Soil (MPBio, USA), according to the
manufacturer’s instructions. e 16S rRNA genes of the genomic DNA were PCR-amplied using the barcoded
universal bacterial primers 27F (5-X-AGA GTT TGA TCM TGG CTC AG-3) and 1492R (5-X-TAC GGY TAC
CTT GTT ACG ACT T-3), where X denotes the 16-nucleotide long barcode. PCR conditions were as follows:
94 °C for 5 min, followed by 32 cycles of 94 °C for 45 s, 55 °C for 45 s, and 72 °C for 1 min, and 1 nal cycle of 72 °C
for 10 min. e PCR products were puried using a PCR purication kit (Solgent, Korea). A pooled composite
was prepared by mixing equal amounts of the puried PCR products and sequenced on a PacBio RS II instrument
using P6-C4 chemistry (Pacic Biosciences, USA) aer SMRTbell adapter ligation at Macrogen (Korea). e
PacBio raw data generated by the PacBio RS II single-molecule real-time (SMRT) sequencing were assembled in
circular consensus sequences (CCSs) using the RS_ReadsOfInsert protocol in the SMRT Analysis Soware (ver.
2.3) with the parameters “minFullPasses 5” and “minPredictedAccuracy 90” aer removing SMRTbell adapters,
and the CCSs were demultiplexed based on their PCR barcodes using the SMRT Portal72. e demultiplexed
sequencing reads were processed using the mothur soware73 as follows: sequences of 1,400–1,600 nucleo-
tides with fewer than two ambiguous nucleotides were selected and chimera sequences were removed using the
UCHIME program of the mothur soware74 against the SILVA Gold reference database (ver. 123). e resulting
high-quality sequences were classied using the naïve Bayesian classier based on the SILVA reference taxonomy,
with an 80% cuto.
Isolation of major bacteria from the enrichment cultures and biodegradation tests. To
isolate major bacteria from the enrichment cultures, the final enrichment cultures were serially diluted in
phosphate-buered saline (137 mM NaCl, 2.7 mM KCl, 10 mM Na2HPO4, and 2 mM KH2PO4, pH 7.2) and spread
onto R2A agar (BD, USA) and the agar plates were incubated at 25 °C. irty colonies from each enrichment were
randomly selected and their 16S rRNA genes were PCR-amplied using the 27F and 1492R primers as described
previously75. e PCR amplicons were double-digested with a mixture of HaeIII and HhaI restriction enzymes
and representative PCR products with unique restriction fragment length polymorphism (RFLP) patterns were
sequenced, as described previously75. e resulting 16S rRNA gene sequences were compared with those of all
reported type strains using the Nucleotide Similarity Search program in the EzTaxon-e server (https://www.
ezbiocloud.net/)76.
e naphthalene-, BTEX-, and hexadecane-biodegrading abilities of the isolates with unique RFLP patterns
were tested in triplicate in sterile 160 ml serum bottles containing 10 ml of minimal salt basal media (MSB)77 sup-
plemented with 30 ppm of naphthalene, 180 ppm of BTEX mixture, and 100 ppm of n-hexadecane, respectively,
as described previously33,75. In brief, cells of the isolates grown in R2A broth at 25 °C were inoculated into the
serum bottles at a density of approximately 107 cells/ml. e serum bottles were sealed with Teon-coated rubber
septa and incubated at 25 °C with shaking (180 rpm) for a week. e incubated media were extracted with meth-
ylene chloride (for BTEX and n-hexadecane) or ethyl acetate (for naphthalene) and the remaining BTEX, naph-
thalene, and n-hexadecane concentrations were measured using gas chromatography, as described previously33,75.
To infer the phylogenetic relationship of strain BN5, which shows naphthalene- and BTEX-biodegrading
ability, the 16S rRNA gene sequences of BN5 and some closely related type strains were aligned using the fast
secondary structure-aware Infernal aligner in the Ribosomal Database Project (RDP) (http://pyro.cme.msu.
edu/spring/align.spr)78, and a phylogenetic tree was constructed using the neighbor-joining algorithm with the
Kimura two-parameter model in the PHYLIP soware79.
Biodegradation of naphthalene and BTEX by strain BN5 in MSB and soil slurry systems. e
biodegradation ability of strain BN5 was evaluated with naphthalene and BTEX in sterile 160 ml serum bottles
containing 10 ml of MSB medium supplemented with naphthalene (30 mg/l) or BTEX mixture (180 mg/l). BN5
cells grown in R2A broth at 30 °C were inoculated at a density of approximately 2 × 107 cells/ml and incubated at
30 °C with shaking (180 rpm). ree bottles per sample were sacriced periodically and the compound concen-
trations were analyzed using HP 7890B gas chromatography (GC) coupled to a ame ionization detector (Agilent,
USA) with an HP-5 column (30 m length, 0.32 µm inner diameter, 0.25 µm lm thickness; J & W Scientic). e
GC oven temperature was held at 40 °C for 0.50 min and increased at a rate of 10 °C per min to 150 °C and then
at a rate of 15 °C per min to a nal temperature of 260 °C, which was held for 5 min. Uninoculated serum bottles
were used as negative controls.
To evaluate the biodegradation ability of strain BN5 with naphthalene and BTEX in soil slurry systems, four
sets of serum bottles (untreated; treated with 1 g/l of NH4Cl and 0.5 g/l of Na3PO4 as nutrients; treated with strain
BN5; and treated with nutrients and strain BN5) were prepared. ree grams of soil obtained from the gasoline
and diesel fuel-contaminated site were transferred into 160 ml serum bottles containing 10 ml of freshwater (col-
lected from the Han River, Korea, July 2017) (supplemented, or unsupplemented with nutrients) with naphtha-
lene (30 mg/l) or BTEX mixture (180 mg/l). BN5 cells were inoculated into the serum bottles at approximately
2 × 107 cells/ml. e serum bottles were incubated at 30 °C with shaking (180 rpm, with periodic inversion) and
three bottles per experimental set were sacriced periodically and the concentrations of naphthalene and BTEX
were analyzed using GC, as described above.
Complete genome sequencing and bioinformatic analyses. For the whole genome sequencing of
strain BN5, genomic DNA was extracted using a Promega Wizard Genomic DNA purication kit (Promega,
USA), according to the manufacturer’s instructions. e genome of strain BN5 was sequenced using a combi-
nation of PacBio RS II SMRT sequencing with a 10-kb library and Illumina Hiseq2500 sequencing at Macrogen
(Korea). De novo assembly of the PacBio sequencing reads was performed through the hierarchical genome
assembly process. Paired-end reads (101-bp) derived from Illumina sequencing were mapped on the complete
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
11
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
genome obtained by the PacBio sequencing for error corrections. e NCBI (http://www.ncbi.nlm.nih.gov/) and
Integrated Microbial Genomes (IMG; https://img.jgi.doe.gov/cgi-bin/er/main.cgi) servers were primarily used
for genome analysis.
The ANI and in silico DDH values between BN5 and its closely related strains were calculated using a
stand-alone soware (http://www.ezbiocloud.net/sw/oat)80 and the server-based genome-to-genome distance
calculator ver. 2.1 (http://ggdc.dsmz.de/distcalc2.php)81, respectively. Circular maps of the chromosomes and
plasmids of strain BN5 were generated using the web-based CGview tool (http://stothard.afns.ualberta.ca/
cgview_server/index.html)82. Clustered regularly interspaced short palindromic repeat (CRISPR) gene sequences
were identied using an online web service (http://crispr.i2bc.paris-saclay.fr/Server/). Genomic islands (GIs) were
predicted using the SIGI-HMM tool with Island Viewer 4 (http://www.pathogenomics.sfu.ca/islandviewer/)83.
e annotation and putative functions of genes associated with the biodegradation of naphthalene, BTEX,
and aliphatic hydrocarbons were bioinformatically analyzed based on data from the NCBI genomic database,
the Integrated Microbial Genomes (IMG) database, UniProt database (http://www.uniprot.org/), and BLAST
searches. e catabolic pathways of naphthalene and each BTEX compound were predicted based on the annota-
tion results of the metabolic genes.
Biodegradation tests of strain BN5 on heptane, nonane, and hexadecane in MSB. e bio-
degradation of heptane, nonane, and n-hexadecane by strain BN5 was evaluated in sterile 160 ml serum bottles
containing 10 ml of MSB medium supplemented with 100 mg/l of each of heptane, nonane, and n-hexadecane,
as described above. BN5 cells were inoculated at a density of approximately 2 × 107 cells/ml and the serum bot-
tles were incubated at 30 °C with shaking (180 rpm) for 5 days. e concentrations of heptane, nonane, and
n-hexadecane were analyzed using GC aer methylene chloride extraction, as described above.
GC/MS detection of naphthalene and BTEX metabolites. e catabolic pathways that were predicted
based on bioinformatic analyses were conrmed through the detection of metabolic intermediates of naphthalene
and BTEX compounds using a GC-MSD, as previously described15,33, with some modication. Triplicate serum
bottles containing 10 ml of MSB media supplemented with naphthalene or each of the individual BTEX com-
pounds (300 ppm) were prepared and incubated at 30 °C with shaking (150 rpm) for 2 days. Triplicate suspen-
sions were combined and acidied with HCl to approximately pH 2.0 and metabolic intermediates were extracted
with 10 ml of ethyl acetate. e ethyl acetate fractions were dried over anhydrous Na2SO4 and concentrated under
an atmosphere of N2 to a volume of 500 µL. Metabolic intermediates were derivatized with 50 µL bis(trimethyl-
silyl) triuoroacetamide (BSTFA) for 30 min and analyzed using a HP 7820 A gas chromatography-5977E mass
spectroscopy detector (GC-MSD) (Agilent, USA) equipped with an HP-5 capillary column. e GC oven tem-
perature was programmed to increase from 50 °C to 260 °C at 5 °C/min and then was held at 260 °C for 5 min. e
peaks were identied by matching query mass spectra to those in a mass spectral reference library.
Phylogenetic analysis of catabolic genes. Homologous amino acid sequences of naphthalene
1,2-dioxygenase large oxygenase component (NagAc), toluene-4-monooxygenase system protein A (TmoA), and
alkane 1-monooxygenase (AlkB) were identied using the UniProt database and maximum-likelihood (ML)
algorithm-based trees showing their phylogenetic relationships were constructed using the MEGA7 program and
the tree topologies were evaluated through bootstrap tests with 1,000 iterations84.
References
1. Naidu, . ecent advances in contaminated site remediation. Water Air Soil Poll 224, 1705 (2013).
2. Namoong, W., Par, J. S. & VanderGheynst, J. S. Bioltration of gasoline vapor by compost media. Environ Pollut 121, 181–187
(2003).
3. Biswas, B., Sarar, B., usmin, . & Naidu, . Bioremediation of PAHs and VOCs: Advances in clay mineral-microbial interaction.
Envi ron Int 85, 168–181 (2015).
4. Cheng, H. F., Hu, E. D. & Hu, Y. A. Impact of mineral micropores on transport and fate of organic contaminants: A review. J Contam
Hydrol 129, 80–90 (2012).
5. Samanta, S. ., Singh, O. V. & Jain, . . Polycyclic aromatic hydrocarbons: environmental pollution and bioremediation. Trends
Biotechnol 20, 243–248 (2002).
6. Dean, B. J. ecent ndings on the genetic toxicology of benzene, toluene, xylenes and phenols. Mutat e s 154, 153–181 (1985).
7. Garr, A. L., Laramore, S. & rebs, W. Toxic eects of oil and dispersant on marine microalgae. Bull Environ Contam Toxicol 93,
654–659 (2014).
8. Harvey, A. N., Snape, I. & Siciliano, S. D. Validating potential toxicity assays to assess petroleum hydrocarbon toxicity in polar soil.
Environ Toxicol Chem 31, 402–407 (2012).
9. Mcee, . H. & White, . e mammalian toxicological hazards of petroleum-derived substances: an overview of the petroleum
industry response to the high production volume challenge program. Int J Toxicol 33, 4S–16S (2014).
10. Ugochuwu, U. C. et al. Exposure riss to polycyclic aromatic hydrocarbons by humans and livestoc (cattle) due to hydrocarbon
spill from petroleum products in Niger-delta wetland. Envir on Int 115, 38–47 (2018).
11. Fuentes, S., Mendez, V., Aguila, P. & Seeger, M. Bioremediation of petroleum hydrocarbons: catabolic genes, microbial communities,
and applications. Appl Microbiol Biotechnol 98, 4781–4794 (2014).
12. NC. In situ bioremediation: When does it wor?. National Academy Press: Washington, DC (1993).
13. Atlas, . M. & Cerniglia, C. E. Bioremediation of petroleum pollutants - diversity and environmental aspects of hydrocarbon
biodegradation. Bioscience 45, 332–338 (1995).
14. Gillespie, I. M. M. & Philp, J. C. Bioremediation, an environmental remediation technology for the bioeconomy. Trends Biotechnol
31, 329–332 (2013).
15. Choi, E. J. et al. Comparative genomic analysis and benzene, toluene, ethylbenzene, and o-, m-, and p-xylene (BTEX) degradation
pathways of Pseudoxanthomonas spadix BD-a59. Appl Environ Microbiol 79, 663–671 (2013).
16. Jiang, B. et al. Bioremediation of petrochemical wastewater containing BTEX compounds by a new immobilized bacterium
Comamonas sp. JB in magnetic gellan gum. Appl Biochem Biotechnol 176, 572–581 (2015).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
12
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
17. Lee, H., Yun, S. Y., Jang, S., im, G. H. & im, J. J. Bioremediation of polycyclic aromatic hydrocarbons in creosote-contaminated
soil by Peniophora incarnata UC8836. Bioremediat J 19, 1–8 (2015).
18. Murinova, S., Dercova, . & Dudasova, H. Degradation of polychlorinated biphenyls (PCBs) by four bacterial isolates obtained from
the PCB-contaminated soil and PCB-contaminated sediment. Int Biodeterior Biodegradation 91, 52–59 (2014).
19. Sun, G. D., Xu, Y., Liu, Y. & Liu, Z. P. Microbial community dynamics of soil mesocosms using Orychophragmus violaceus combined wit h
hodococcus ruber Em1 for bioremediation of highly PAH-contaminated soil. Appl Microbiol Biotechnol 98, 10243–10253 (2014).
20. avamani, P., Megharaj, M. & Naidu, . Bioremediation of high molecular weight polyaromatic hydrocarbons co-contaminated
with metals in liquid and soil slurries by metal tolerant PAHs degrading bacterial consortium. Biodegradation 23, 823–835 (2012).
21. Tyagi, M., da Fonseca, M. M. . & de Carvalho, C. C. C. . Bioaugmentation and biostimulation strategies to improve the
eectiveness of bioremediation processes. Biodegradation 22, 231–241 (2011).
22. ayu, S., arpouzas, D. G. & Singh, B. . Emerging technologies in bioremediation: constraints and opportunities. Biodegradation
23, 917–926 (2012).
23. vanVeen, J. A., van Overbee, L. S. & van Elsas, J. D. Fate and activity of microorganisms introduced into soil. Microbiol Mol Biol ev
61, 121–135 (1997).
24. Andreolli, M., Lampis, S., Zenaro, E., Salinoja-Salonen, M. & Vallini, G. Burholderia f ungorum DBT1: a promising bacterial strain
for bioremediation of PAHs-contaminated soils. FEMS Microbiol Lett 319, 11–18 (2011).
25. M’rassi, A. G., Bensalah, F., Gury, J. & Duran, . Isolation and characterization of dierent bacterial strains for bioremediation of
n-alanes and polycyclic aromatic hydrocarbons. Environ Sci Pollut es 22, 15332–15346 (2015).
26. Hedlund, B. P. & Staley, J. T. Isolation and characterization of Pseudoalteromonas strains with divergent polycyclic aromatic
hydrocarbon catabolic properties. Environ Microbiol 8, 178–182 (2006).
27. Jeon, C. O. et al. Discovery of a bacterium, with distinctive dioxygenase, that is responsible for in situ biodegradation in
contaminated sediment. Proc Natl Acad Sci USA 100, 13591–13596 (2003).
28. asai, Y., ishira, H. & Harayama, S. Bacteria belonging to the genus Cycloclasticus play a primary role in the degradation of
aromatic hydrocarbons released in a marine environment. Appl Environ Microbiol 68, 5625–5633 (2002).
29. Sohn, J. H., won, . ., ang, J. H., Jung, H. B. & im, S. J. Novosphingobium pentaromativorans sp. nov., a high-molecular-mass polycyclic
aromatic hydrocarbon-degrading bacterium isolated from estuarine sediment. Int J Syst Evol Microbiol 54, 1483–1487 (2004).
30. Uhli, O. et al. Identication of bacteria utilizing biphenyl, benzoate, and naphthalene in long-term contaminated soil. PLoS ONE 7,
e40653 (2012).
31. Duan, T. H. & Adrian, L. Enrichment of hexachlorobenzene and 1,3,5-trichlorobenzene transforming bacteria from sediments in
Germany and Vietnam. Biodegradation 24, 513–520 (2013).
32. Jacques, . J. S., Oee, B. C., Bento, F. M., Peralba, M. C. . & Camargo, F. A. O. Improved enrichment and isolation of polycyclic
aromatic hydrocarbons (PAH)-degrading microorganisms in soil using anthracene as a model PAH. Curr Microbiol 58, 628–634
(2009).
33. Jin, H. M., im, J. M., Lee, H. J., Madsen, E. L. & Jeon, C. O. Alteromonas as a ey agent of polycyclic aromatic hydrocarbon
biodegradation in crude oil-contaminated coastal sediment. Environ Sci Technol 46, 7731–7740 (2012).
34. im, J. M. & Jeon, C. O. Isolation and characterization of a new benzene, toluene, and ethylbenzene degrading bacterium,
Acinetobacter sp B113. Curr Microbiol 58, 70–75 (2009).
35. Tai, H., Syutsubo, ., Mattison, . G. & Harayama, S. Identication and characterization of o-xylene-degrading hodococcus spp.
which were dominant species in the remediation of o-xylene-contaminated soils. Biodegradation 18, 17–26 (2007).
36. Andreoni, V. & Gianfreda, L. Bioremediation and monitoring of aromatic-polluted habitats. Appl Microbiol Biotechnol 76, 287–308
(2007).
37. Oberoi, A. S., Philip, L. & Bhallamudi, S. M. Biodegradation of various aromatic compounds by enriched bacterial cultures: part
A-monocyclic and polycyclic aromatic hydrocarbons. Appl Biochem Biotechnol 176, 1870–1888 (2015).
38. Auret, M., Labbe, D., ouand, G., Greer, C. W. & Fayolle-Guichard, F. Degradation of a mixture of hydrocarbons, gasoline, and
diesel oil additives by hodococcus aetherivorans and hodococcus wratislaviensis. Appl Environ Microbiol 75, 7774–7782 (2009).
39. Perez-Pantoja, D. et al. Genomic analysis of the potential for aromatic compounds biodegradation in Burholderiales. Environ
Microbiol 14, 1091–1117 (2012).
40. Mitter, B. et al. Comparative genome analysis of Burholderia phytormans PsJN reveals a wide spectrum of endophytic lifestyles
based on interaction strategies with host plants. Front Plant Sci 4, 120 (2013).
41. Weilharter, A. et al. Complete genome sequence of the plant growth-promoting endophyte Burholderia phytormans strain PsJN.
J Bacteriol 193, 3383–3384 (2011).
42. im, M., Oh, H.-S., Par, S.-C. & Chun, J. Towards a taxonomic coherence between average nucleotide identity and 16S rNA gene
sequence similarity for species demarcation of proaryotes. Int J Syst Evol Microbiol 64, 346–351 (2014).
43. osselló-Móra, . & Amann, . Past and future species denitions for Bacteria and Archaea. Syst Appl Microbiol 38, 209–216 (2015).
44. Lee, Y. & Jeon, C. O. Paraburholderia aromaticivorans sp. nov., an aromatic hydrocarbon-degrading bacterium, isolated from
gasoline-contaminated soil. Int J Syst Evol Microbiol 68, 1251–1257 (2018).
45. Peng, . H. et al. Microbial biodegradation of polyaromatic hydrocarbons. FEMS Microbiol ev 32, 927–955 (2008).
46. Math, . . et al. Comparative genomics reveals adaptation by Alteromonas sp. SN2 to marine tidal-at conditions: cold tolerance
and aromatic hydrocarbon metabolism. Plos One 7, e35784 (2012).
47. Par, M. J. et al. Molecular and biochemical characterization of 3-hydroxybenzoate 6-hydroxylase from Polaromonas
naphthalenivorans CJ2. Appl Environ Microbiol 73, 5146–5152 (2007).
48. Jeon, C. O., Par, M., o, H. S., Par, W. & Madsen, E. L. e naphthalene catabolic (nag) genes of Polaromonas naphthalenivorans
CJ2: evolutionary implications for two gene clusters and novel regulatory control. Appl Environ Microbiol 72, 1086–1095 (2006).
49. Zhou, N. Y., Fuenmayor, S. L. & Williams, P. A. nag genes of alstonia (formerly Pseudomonas) sp strain U2 encoding enzymes for
gentisate catabolism. J Bacteriol 183, 700–708 (2001).
50. Tao, Y., Fishman, A., Bentley, W. E. & Wood, T. . Oxidation of benzene to phenol, catechol, and 1,2,3-trihydroxybenzene by toluene
4-monooxygenase of Pseudomonas mendocina 1 and toluene 3-monooxygenase of alstonia picettii PO1. Appl Environ
Microbiol 70, 3814–3820 (2004).
51. Zamanian, M. & Mason, J. . Benzene dioxygenase in Pseudomonas putida - subunit composition and immuno-cross-reactivity with
other aromatic dioxygenases. Biochem J 244, 611–616 (1987).
52. Olajire, A. & Essien, J. Aerobic degradation of petroleum components by microbial consortia. J Pet Environ Biotechnol 5, 5 (2014).
53. Cao, B., Nagarajan, . & Loh, . C. Biodegradation of aromatic compounds: current status and opportunities for biomolecular
approaches. Appl Microbiol Biotechnol 85, 207–228 (2009).
54. Gibson, D. T., Mahadevan, V. & Davey, J. F. Bacterial metabolism of para- and meta-xylene: oxidation of the aromatic ring. J Bacteriol
119, 930–936 (1974).
55. Yu, H., im, B. J. & ittmann, B. E. e roles of intermediates in biodegradation of benzene, toluene, and p-xylene by Pseudomonas
putida F1. Biodegradation 12, 455–463 (2001).
56. Gibson, D. T., Gschwendt, B., Yeh, W. . & obal, V. M. Initial reactions in the oxidation of ethylbenzene by Pseudomonas putida.
Biochemistry 12, 1520–1528 (1973).
57. Lee, . & Gibson, D. T. Toluene and ethylbenzene oxidation by puried naphthalene dioxygenase from Pseudomonas sp strain NCIB
9816-4. Appl Environ Microbiol 62, 3101–3106 (1996).
Content courtesy of Springer Nature, terms of use apply. Rights reserved
www.nature.com/scientificreports/
13
SCIENTIFIC REPORTS | (2019) 9:860 | https://doi.org/10.1038/s41598-018-36165-x
58. Gao, J., Ellis, L. B. & Wacett, L. P. e University of Minnesota biocatalysis/biodegradation database: improving public access.
Nucleic Acids es 38, D488–491 (2010).
59. Mallic, S., Charaborty, J. & Dutta, T. . ole of oxygenases in guiding diverse metabolic pathways in the bacterial degradation of
low-molecular-weight polycyclic aromatic hydrocarbons: A review. Crit ev Microbiol 37, 64–90 (2011).
60. Parales, . E., Parales, J. V., Pelletier, D. A. & Ditty, J. L. Diversity of microbial toluene degradation pathways. Adv Appl Microbiol 64,
1–73 (2008).
61. Seo, J. S., eum, Y. S. & Li, Q. X. Bacterial degradation of aromatic compounds. Int J Environ es Public Health 6, 278–309 (2009).
62. Stotzy, G. Microbial respiration. C.A. Blac (Ed.), Methods of Soil Analysis. SSSA, Madison, 550–1572 (1965).
63. Johnsen, A. ., Wic, L. Y. & Harms, H. Principles of microbial PAH-degradation in soil. Environ Pollut 133, 71–84 (2005).
64. Mason, O. U. et al. Metagenomics reveals sediment microbial community response to Deepwater Horizon oil spill. ISME J 8,
1464–1475 (2014).
65. Li, J. B. et al. Biodegradation of phenanthrene in polycyclic aromatic hydrocarbon-contaminated wastewater revealed by coupling
cultivation-dependent and -independent approaches. Environ Sci Technol 51, 3391–3401 (2017).
66. Gerdes, B., Brinmeyer, ., Diecmann, G. & Helme, E. Inuence of crude oil on changes of bacterial communities in Arctic sea-
ice. FEMS Microbiol Ecol 53, 129–139 (2005).
67. Aagot, N., Nybroe, O., Nielsen, P. & Johnsen, . An altered Pseudomonas diversity is recovered from soil by using nutrient-poor
Pseudomonas-selective soil extract media. Appl Environ Microbiol 67, 5233–5239 (2001).
68. McCaig, A. E., Grayston, S. J., Prosser, J. I. & Glover, L. A. Impact of cultivation on characterisation of species composition of soil
bacterial communities. FEMS Microbiol Ecol 35, 37–48 (2001).
69. Sorheim, ., Torsvi, V. L. & Gosoyr, J. Phenotypical divergences between populations of soil bacteria isolated on dierent media.
Microbiol Ecol 17, 181–192 (1989).
70. Singh, B. . & Waler, A. Microbial degradation of organophosphorus compounds. FEMS Microbiol ev 30, 428–471 (2006).
71. Afzal, M., han, S., Iqbal, S., Mirza, M. S. & han, Q. M. Inoculation method aects colonization and activity of Burholderia
phytormans PsJN during phytoremediation of diesel-contaminated soil. Int Biodeterior Biodegradation 85, 331–336 (2013).
72. Fichot, E. B. & Norman, . S. Microbial phylogenetic proling with the Pacic Biosciences sequencing platform. Microbiome 1, 10
(2013).
73. Schloss, P. D. et al. Introducing mothur: open-source, platform-independent, community-supported soware for describing and
comparing microbial communities. Appl Environ Microbiol 75, 7537–7541 (2009).
74. Edgar, . C., Haas, B. J., Clemente, J. C., Quince, C. & night, . UCHIME improves sensitivity and speed of chimera detection.
Bioinformatics 27, 2194–2200 (2011).
75. im, J. M. et al. Inuence of soil components on the biodegradation of benzene, toluene, ethylbenzene, and o-, m-, and p-xylenes by
the newly isolated bacterium Pseudoxanthomonas spadix BD-a59. Appl Environ Microbiol 74, 7313–7320 (2008).
76. Yoon, S. H. et al. Introducing EzBioCloud: a taxonomically united database of 16S rNA gene sequences and whole-genome
assemblies. Int J Syst Evol Microbiol 67, 1613–1617 (2017).
77. Stanier, . Y., Palleroni, N. J. & Doudoro, M. e aerobic pseudomonads: a taxonomic study. J Gen Microbiol 43, 159–271 (1966).
78. Nawroci, E. P. & Eddy, S. . Query-dependent banding (QDB) for faster NA similarity searches. PLoS Comput Biol 3, e56 (2007).
79. Felsenstein, J. PHYLIP (phylogeny inference pacage), version 3.6a. Seattle: Department of genetics, University of Washington,
Seattle, WA, USA (2002).
80. Lee, I., im, Y. O., Par, S.-C. & Chun, J. OrthoANI: an improved algorithm and soware for calculating average nucleotide identity.
Int J Syst Evol Microbiol 66, 1100–1103 (2016).
81. Meier-oltho, J. P., Auch, A. F., len, H.-P. & Göer, M. Genome sequence-based species delimitation with condence intervals
and improved distance functions. BMC bioinformatics 14, 60 (2013).
82. Stothard, P. & Wishart, D. S. Circular genome visualization and exploration using CGView. Bioinformatics 21, 537–539 (2005).
83. Langille, M. G. I. & Brinman, F. S. L. IslandViewer: an integrated interface for computational identication and visualization of
genomic islands. Bioinformatics 25, 664–665 (2009).
84. umar, S., Stecher, G. & Tamura, . MEGA7: molecular evolutionary genetics analysis version 7.0 for bigger datasets. Mol Biol Evol
33, 1870–1874 (2016).
Acknowledgements
is work was supported by the National Research Foundation (2017R1A2B4004888, 2018R1A5A1025077) of
the Ministry of Science, ICT and Future Planning, Republic of Korea.
Author Contributions
C.O.J. conceived the project. Yunho Lee and Yunhee Lee performed the experiments and conducted the
bioinformatics analysis. Yunho Lee and C.O.J. wrote the paper. All authors read and approved the final
manuscript.
Additional Information
Supplementary information accompanies this paper at https://doi.org/10.1038/s41598-018-36165-x.
Competing Interests: e authors declare no competing interests.
Publisher’s note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional aliations.
Open Access This article is licensed under a Creative Commons Attribution 4.0 International
License, which permits use, sharing, adaptation, distribution and reproduction in any medium or
format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Cre-
ative Commons license, and indicate if changes were made. e images or other third party material in this
article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the
material. If material is not included in the article’s Creative Commons license and your intended use is not per-
mitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.
© e Author(s) 2019
Content courtesy of Springer Nature, terms of use apply. Rights reserved
1.
2.
3.
4.
5.
6.
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for small-
scale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
use bots or other automated methods to access the content or redirect messages
override any security feature or exclusionary protocol; or
share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com

Supplementary resource (1)

... The Paraburkholderia strain was demonstrated to belong to P. bryophila species and could use a broad range of monosaccharide, sugar alcohols, organic acids and amino acids. This is in line with the high metabolic versatility known for several Paraburkholderia strains (Ivaldi et al., 2023;Lee et al., 2019). Members of Paraburkholderia bryophila that were previously isolated from soils and mosses were reported to have antimicrobial properties and phosphate solubilization capacity (Hsu et al., 2018;Vandamme et al., 2007). ...
Article
Full-text available
Chlorobenzene (CB), extensively used in industrial processes, has emerged as a significant contaminant in soil and groundwater. The eco-friendly and cost-effective microbial remediation has been increasingly favored to address this environmental challenge. In this study, a degrading bacterium was isolated from CB-contaminated soil at a pesticide plant, identified as Pandoraea sp. XJJ-1 (CCTCC M 2021057). This strain completely degraded 100 mg·L⁻¹ CB and showed extensive degradability across a range of pH (5.0–9.0), temperature (10–37 °C), and CB concentrations (100–600 mg·L⁻¹). Notably, the degradation efficiency was 85.2% at 15 °C, and the strain could also degrade six other aromatic hydrocarbons, including benzene, toluene, ethylbenzene, and xylene (o-, m-, p-). The metabolic pathway of CB was inferred using ultraperformance liquid chromatography, gas chromatography-mass spectrometry, and genomic analysis. In strain XJJ-1, CB was metabolized to o-chlorophenol and 3-chloroxychol by CB monooxygenase, followed by ortho-cleavage by the action of 3-chlorocatechol 1,2-dioxygenase. Moreover, the presence of the chlorobenzene monooxygenation pathway metabolism in strain XJJ-1 is reported for the first time in Pandoraea. As a bacterium with low-temperature resistance and composite pollutant degradation capacity, strain XJJ-1 has the potential application prospects in the in-situ bioremediation of CB-contaminated sites.
Article
Full-text available
The biodegradation of diverse organic compounds plays a pivotal role in environmental sustainability and waste management. In recent years, genetically engineered microbial cells have emerged as a promising tool to enhance biodegradation capabilities. This study presents an innovative approach to produce amino acids and nucleic acids by manipulating microbial genomes, shedding light on their relevance to biodegradation. Through the power of genetic engineering, microbial cells can be harnessed to efficiently degrade a wide range of pollutants, contributing to a cleaner and healthier environment. This study explores the transformative potential of genetic engineering in unlocking the biodegradative capacity of microbial cells and highlights its significance in the production of amino acids and nucleic acids. By leveraging these advancements, we can pave the way towards sustainable waste management and environmental remediation, ensuring a greener and more sustainable future.
Article
Full-text available
Phenolic root exudates (PREs) secreted by wetland plants facilitate the accumulation of iron in the rhizosphere, potentially providing the essential active iron required for the generation of enzymes that degrade polycyclic aromatic hydrocarbon, thereby enhancing their biodegradation. However, the underlying mechanisms involved are yet to be elucidated. This study focuses on phenanthrene (PHE), a typical polycyclic aromatic hydrocarbon pollutant, utilizing representative PREs from wetland plants, including p-hydroxybenzoic, p-coumaric, caffeic, and ferulic acids. Using hydroponic experiments, 16S rRNA sequencing, and multiple characterization techniques, we aimed to elucidate the interaction mechanism between the accelerated degradation of PHE and the formation of rhizosphere biofilm/iron plaque influenced by PREs. Although all four types of PREs altered the biofilm composition and promoted the formation of iron plaque on the root surface, only caffeic acid, possessing a similar structure to the intermediate metabolite of PHE (catechol), could accelerate the PHE degradation rate. Caffeic acid, notable for its catechol structure, plays a significant role in enhancing PHE degradation through two main mechanisms: (a) it directly boosts PHE co-metabolism by fostering the growth of PHE-degrading bacteria, specifically Burkholderiaceae, and by facilitating the production of the key metabolic enzyme catechol 1,2-dioxygenase (C12O) and (b) it indirectly supports PHE biodegradation by promoting iron plaque formation on root surfaces, thereby enriching free iron for efficient microbial synthesis of C12O, a crucial factor in PHE decomposition.
Article
Paraburkholderia terrae strain KU-46 has been studied for its capability to degrade 2,4-dinitrophenol. Here, we present the complete 10,833,180bp genome of this microorganism, comprising five circular chromosomes housing 9,797 protein-coding sequences. The genes responsible for 2,4-dinitrophenol and 4-nitrophenol degradation are located on chromosome 2.
Article
Full-text available
The overall impact of a crude oil spill into a pristine freshwater environment in Canada is largely unknown. To evaluate the impact on the native microbial community, a large-scale in situ model experimental spill was conducted to assess the potential role of the natural community to attenuate hydrocarbons. A small volume of conventional heavy crude oil (CHV) was introduced within contained mesocosm enclosures deployed on the shoreline of a freshwater lake. The oil was left to interact with the shoreline for 72 h and then free-floating oil was recovered using common oil spill response methods (i.e. freshwater flushing and capture on oleophilic absorptive media). Residual PAH concentrations returned to near pre-oiling concentrations within 2 months, while the microbial community composition across the water, soil, and sediment matrices of the enclosed oligotrophic freshwater ecosystems did not shift significantly over this period. Metagenomic analysis revealed key polycyclic aromatic and alkane degradation mechanisms also did not change in their relative abundance over the monitoring period. These trends suggest that for small spills (<2 L of oil per 15 m2 of surface freshwater), physical oil recovery reduces PAH concentrations to levels tolerated by the native microbial community. Additionally, the native microbial community present in the monitored pristine freshwater ecosystem possesses the appropriate hydrocarbon degradation mechanisms without prior challenge by hydrocarbon substrates. This study corroborated trends found previously (Kharey et al. 2024) toward freshwater hydrocarbon degradation in an environmentally relevant scale and conditions on the tolerance of residual hydrocarbons in situ.
Chapter
Pollutants can exert numerous adverse impacts on both human health and the environment. To minimize pollutants and their effects, it is important to reduce emissions, practice proper waste management, and use cleaner technologies. Polluted environments can be cleaned using bioremediation, a more cost-effective and environmentally friendly method than traditional methods. Several bioremediation techniques are available, including biostimulation, which stimulates the growth of naturally occurring microorganisms, and bioaugmentation, which enhances natural biodegradation by adding specific microorganisms or enzymes to the contaminated environment. Chemicals, pesticides, petroleum hydrocarbons, heavy metals, and petroleum hydrocarbons have all been successfully cleaned up using bioremediation. The use of genetically engineered microbes (GEMs) for environmental remediation involves modifying microorganisms to degrade or remove pollutants from contaminated environments. Although GEMs show promise as a solution to environmental pollution, there are also concerns about their potential unintended consequences, such as the spread of modified genes to other organisms. To implement GEMs on a large scale, it is important to evaluate the potential risks and benefits thoroughly. Treatment with microbes is a less costly and more effective method of removing pollutants from polluted environments. This chapter discussed challenges in designing multi-functional genetic engineering microorganisms for simultaneous pollutant removal from soil contaminated by GEMs and recent advances in bioremediation techniques and approaches.
Article
Full-text available
Paraburkholderia sp. strain 22B1P utilizes 3-chlorobenzoate as a carbon source. Complete genome sequencing of strain 22B1P revealed two chromosomes and two plasmids. The genes involved in the conversion of 3-chlorobenzoate to 3-chlorocatechol and those involved in the conversion of 3-chlorocatechol to 3-oxoadipate were located on chromosomes 2 and 1, respectively.
Article
Full-text available
Currently, the pollution of the Caspian Sea by the oil industry is one of the highest problems in this area. Critically endangered species inhabit this sea, such as sturgeons, whose ecological value is incalculable. Thus, we aimed to evaluate the level of contamination of aliphatic hydrocarbons of petroleum and its relation with several toxic elements directly on sturgeons spines. A total of 40 adult starry sturgeons (Acipenser stellatus) were obtained within a repopulation programme in the northern and southern coastal waters of the Caspian Sea. The marginal pectoral fin was extracted from each fish to determine aliphatic hydrocarbons, arsenic, cadmium, mercury, nickel, lead, and vanadium. Subsequently, the sturgeons were released. Clearly, the presence of hydrocarbons was evidenced in all the sampled areas finding higher concentrations in the northern areas (N1 = 1.35 ± 0.4; N2 = 1.65 ± 0.46; N3 = 1.27 ± 0.40; S1 = 0.61 ± 0.22; S2 = 0.85 ± 0.43 mg/kg). Furthermore, to a greater or lesser extent, some toxic elements, mainly Hg and As, have been linked to aliphatic hydrocarbons.
Article
Full-text available
We present the latest version of the Molecular Evolutionary Genetics Analysis (MEGA) software, which contains many sophisticated methods and tools for phylogenomics and phylomedicine. In this major upgrade, MEGA has been optimized for use on 64-bit computing systems for analyzing bigger datasets. Researchers can now explore and analyze tens of thousands of sequences in MEGA. The new version also provides an advanced wizard for building timetrees and includes a new functionality to automatically predict gene duplication events in gene family trees. The 64-bit MEGA is made available in two interfaces: graphical and command line. The graphical user interface (GUI) is a native Microsoft Windows application that can also be used on Mac OSX. The command line MEGA is available as native applications for Windows, Linux, and Mac OSX. They are intended for use in high-throughput and scripted analysis. Both versions are available from www.megasoftware.net free of charge.
Article
Full-text available
The recent advent of DNA sequencing technologies facilitates the use of genome sequencing data that provide means for more informative and precise classification and identification of Bacteria and Archaea. Because the current species definition is based on the comparison of genome sequences between type and other strains in a given species, building a genome database with correct taxonomic information is a premium need to enhance our efforts in exploring prokaryotic diversity and discovering new species as well as for routine identifications. Here we introduce an integrated database, called EzBioCloud, that holds the taxonomic hierarchy of Bacteria and Archaea that are represented by quality-controlled 16S rRNA gene and genome sequences. Whole genome assemblies in the NCBI Assembly Database were screened for low quality and subjected to a composite identification bioinformatics pipeline that employs gene-based searches followed by the calculation of average nucleotide identity. As a result, the database is made of 61,700 species/phylotypes including 13,132 with validly published names, and 62,362 whole genome assemblies that were taxonomically identified at the genus, species and subspecies levels. Genomic properties, such as genome size and GC content, and the occurrence in human microbiome data were calculated for each genus or higher taxa. This united database of taxonomy, 16S rRNA gene and genome sequences, with accompanied bioinformatics tools, should accelerate genome-based classification and identification of Bacteria and Archaea. The database and related search tools are available at http://www.ezbiocloud.net/.
Article
Introduced microorganisms are potentially powerful agents for manipulation of processes and/or components in soil. Fields of application include enhancement of crop growth, protection of crops against plant-pathogenic organisms, stimulation of biodegradation of xenobiotic compounds (bioaugmentation), and improvement of soil structure. Inoculation of soils has already been applied for decades, but it has often yielded inconsistent or disappointing results. This is caused mainly by a commonly observed rapid decline in inoculant population activity following introduction into soil, i.e., a decline of the numbers of inoculant cells and/or a decline of the (average) activity per cell. In this review, we discuss the available information on the effects of key factors that determine the fate and activity of microorganisms introduced into soil, with emphasis on bacteria. The factors addressed include the physiological status of the inoculant cells, the biotic and abiotic interactions in soil, soil properties, and substrate availability. Finally, we address the possibilities available to effectively manipulate the fate and activity of introduced microorganisms in relation to the main areas of their application.
Article
A Gram-stain-negative, facultatively aerobic, aromatic hydrocarbon-degrading bacterium, designated strain BN5T, was isolated from gasoline-contaminated soil. Cells were motile and slightly curved rods with a single flagellum showing catalase and oxidase activities. Growth was observed at 20-37 °C (optimum, 25-30 °C), pH 3-7 (optimum, pH 5-6) and 0-2 % NaCl (optimum, 0 %). Ubiquinone-8 was the predominant respiratory quinone. The major fatty acids were C16 : 0, cyclo-C19 : 0ω8c and summed feature 8 (comprising C18 : 1ω7c and/or C18 : 1ω6c). Diphosphatidylglycerol, phosphatidylglycerol, phosphatidylethanolamine, an unidentified phosphoamino lipid, three unidentified amino lipids and eight unidentified lipids were the identified polar lipids. The DNA G+C content was 62.93 mol%. Phylogenetic analysis based on 16S rRNA gene sequences revealed that strain BN5Tformed a phylogenic lineage with members of the genus Paraburkholderia and showed the highest 16S rRNA gene sequence similarities to Paraburkholderia phytofirmans PsJNT(99.4 %), Paraburkholderia dipogonis DL7T(98.8 %) and Paraburkholderia insulsa PNG-AprilT(98.8 %). The average nucleotide identity and in silico DNA-DNA hybridization (DDH) values between strain BN5Tand P. phytofirmans PsJNTwere 88.5 and 36.5 %, respectively. The DDH values for strain BN5Twith P. dipogonis LMG 28415Tand P. insulsa DSM 28142Twere 41.0±4.9 % (reciprocal, 33.0±4.3 %) and 47.1±6.6 % (reciprocal, 51.7±5.4 %), respectively. Based on its physiological, chemotaxonomic and phylogenetic features, we conclude that strain BN5Tis a novel species of the genus Paraburkholderia, for which the name Paraburkholderiaaromaticivorans sp. nov. is proposed. The type strain is BN5T(=KACC 19419T=JCM 32303T).
Article
The indigenous microorganisms responsible for degrading phenanthrene (PHE) in polycyclic aromatic hydrocarbons (PAHs)-contaminated wastewater were identified by DNA-based stable isotope probing (DNA-SIP). In addition to the well-known PHE degraders Acinetobacter and Sphingobium, Kouleothrix and Sandaracinobacter were found, for the first time, to be directly responsible for indigenous PHE biodegradation. Additionally, a novel PHE degrader, Acinetobacter tandoii sp. LJ-5, was identified by DNA-SIP and direct cultivation. This is the first report and reference to A. tandoii involved in the bioremediation of PAHs-contaminated water. A PAH-RHDɑ gene involved in PHE metabolism was detected in the heavy fraction of 13C treatment, but the amplification of PAH-RHDɑ gene failed in A. tandoii LJ-5. Instead, the strain contained catechol 1,2-dioxygenase and the alpha/beta subunits of protocatechuate 3,4-dioxygenase, indicating use of the β-ketoadipate pathway to degrade PHE and related aromatic compounds. These findings add to our current knowledge on microorganisms degrading PHE by combining cultivation-dependent and cultivation-independent approaches and provide deeper insight into the diversity of indigenous PHE-degrading communities.