Skoči na glavni sadržaj

Izvorni znanstveni članak

https://doi.org/10.17113/ftb.59.04.21.7253

Identifikacija i karakterizacija nove plazmidno kodirane višebakrene oksidaze slične lakazama iz bakterije Ochrobactrum sp. BF15, izolirane iz sustava za biološko pročišćavanje poljoprivrednih otpadnih voda

María Carla Martini orcid id orcid.org/0000-0003-0411-7401 ; IBBM – Institute of Biotechnology and Molecular Biology, CONICET - Department of Biological Sciences, Faculty of Exact Sciences, National University of La Plata, Calles 47 y 115 (1900) La Plata, Argentina
Francesca Berini orcid id orcid.org/0000-0003-1913-9929 ; Department of Biotechnology and Life Sciences, University of Insubria, via J.H. Dunant 3, 21100 Varese, Italy
Luka Ausec orcid id orcid.org/0000-0003-3155-0999 ; Department of Food Science and Technology, Biotechnical Faculty, University of Ljubljana, Večna pot 111, 1000 Ljubljana, Slovenia
Carmine Casciello orcid id orcid.org/0000-0002-2966-5124 ; Department of Biotechnology and Life Sciences, University of Insubria, via J.H. Dunant 3, 21100 Varese, Italy
Carolina Vacca orcid id orcid.org/0000-0003-4532-6330 ; IBBM – Institute of Biotechnology and Molecular Biology, CONICET - Department of Biological Sciences, Faculty of Exact Sciences, National University of La Plata, Calles 47 y 115 (1900) La Plata, Argentina
Mariano Pistorio orcid id orcid.org/0000-0001-7931-8721 ; IBBM – Institute of Biotechnology and Molecular Biology, CONICET - Department of Biological Sciences, Faculty of Exact Sciences, National University of La Plata, Calles 47 y 115 (1900) La Plata, Argentina
Antonio Lagares orcid id orcid.org/0000-0003-2644-8324 ; IBBM – Institute of Biotechnology and Molecular Biology, CONICET - Department of Biological Sciences, Faculty of Exact Sciences, National University of La Plata, Calles 47 y 115 (1900) La Plata, Argentina
Ines Mandic-Mulec orcid id orcid.org/0000-0002-7629-3255 ; Department of Food Science and Technology, Biotechnical Faculty, University of Ljubljana, Večna pot 111, 1000 Ljubljana, Slovenia
Flavia Marinelli orcid id orcid.org/0000-0001-9195-6777 ; Department of Biotechnology and Life Sciences, University of Insubria, via J.H. Dunant 3, 21100 Varese, Italy
María Florencia Del Papa orcid id orcid.org/0000-0003-0088-6756 ; IBBM – Institute of Biotechnology and Molecular Biology, CONICET - Department of Biological Sciences, Faculty of Exact Sciences, National University of La Plata, Calles 47 y 115 (1900) La Plata, Argentina


Puni tekst: engleski pdf 966 Kb

str. 519-529

preuzimanja: 186

citiraj

Preuzmi JATS datoteku

Prilozi: FTB-59-519_S1.pdf


Sažetak

Pozadina istraživanja. Posljednjih su desetljeća lakaze (p-difenol:dioksigen oksidoreduktaze; EC 1.10.3.2) privukle pažnju znanstvenika zbog raznolikosti moguće primjene u biotehnologiji i industriji. Lakaze mogu oksidirati različite organske i anorganske spojeve u biokatalizatore prikladne za biotehnološke procese. Iako se u industriji najčešće koriste lakaze iz plijesni, lakaze iz bakterija pokazale su veliki potencijal primjene jer djeluju na više raznih supstrata u različitim uvjetima. Svrha je ovoga rada bila opisati plazmidno kodiranu višebakrenu oksidazu sličnu lakazama iz bakterije Ochrobactrum sp. BF15, prethodno izolirane iz zagađenog tla.
Eksperimentalni pristup. Za identifikaciju novih gena za oksidazu sličnu lakazi u bakteriji Ochrobactrum sp. BF15 koristili smo in silico profil skrivenih Markovljevih modela. Da bismo okarakterizirali lakazu, proveli smo heterolognu ekspresiju gena u bakteriji Escherichia coli, pročistili enzim i mjerili njegovu aktivnost u supstratima karakterističnim za lakazu.
Rezultati i zaključci. Dobiveni profili skrivenih Markovljevih modela omogućili su identifikaciju nove oksidaze slične lakazi, koju smo nazvali Lac80. In silico analiza nove Lac80 pokazala je prisutnost triju konzerviranih domena bakrene oksidaze karakterističnih za lakaze s tri domene. Uspješno smo proveli heterolognu ekspresiju Lac80 u E. coli, čime nam je omogućeno pročišćavanje proteina za daljnju analizu aktivnosti. Od trinaest ispitanih supstrata tipičnih za lakazu, Lac80 je imao slabiju aktivnost u 2,2’-azino-bis(3-etilbenzotiazolin-6-sulfonskoj kiselini) (ABTS), pirokateholu, pirogalolu i vanilinskoj kiselini, te veću aktivnost u 2,6-dimetoksifenolu.
Novina i znanstveni doprinos. Dobiveni rezultati pokazuju da Lac80 može imati obećavajuću primjenu u industriji. U ovom je radu prikazan značaj lakaza iz bakterija te je naglašena važnost plazmida izoliranih iz okoliša kao vrijednih izvora novih gena koji kodiraju enzime za primjenu u biotehnološkim procesima.

Ključne riječi

višebakrene oksidaze slične lakazama; Ochrobactrum; sustav za biološko pročišćavanje; plazmid; biološka razgradnja; heterologna ekspresija

Hrčak ID:

270630

URI

https://hrcak.srce.hr/270630

Datum izdavanja:

22.12.2021.

Podaci na drugim jezicima: engleski

Posjeta: 902 *




INTRODUCTION

Enzymes traditionally known as laccases are multicopper oxidases that catalyse the oxidation of a wide range of substrates and the simultaneous reduction of molecular oxygen to water. Due to their ability to oxidize a wide variety of phenolic and non-phenolic compounds, they have been extensively used in biotechnological processes as biocatalysts (1,2). They were recently renamed ’laccase-like multicopper oxidases’ (LMCOs), following the nomenclature revision of this enormously diverse group of enzymes (3). LMCOs are produced by a wide range of organisms such as bacteria, higher plants, insects and fungi, the fungi being the more broadly studied and characterized, in part due to the development of efficient expression systems in yeasts that have potentiated their industrial use, particularly in the textile and food industry (4-7). However, much attention has been given more recently to prokaryote-sourced LMCOs. Bacterial laccases usually display higher thermal and alkaline pH stability than their eukaryotic counterparts and are active on a broad range of substrates, unlike most fungal laccases (3,8), making them suitable for different industrial processes. A computational study analyzing over 2000 bacterial genomes predicted the presence of laccase-like encoding genes in 36% of the screened organisms (9) including autotrophic, alkalophilic and even anaerobic bacteria (10), highlighting the extraordinary potential of bacteria as a source of novel LMCOs and pointing out the need to continue exploring the genome of these microorganisms.

Multiple approaches have been applied for the identification of novel LMCOs in bacteria, including classical functional- and/or sequence-based screenings of culturable microbiota (11), or metagenomic analyses (10,12). In addition, the generation of chimeric laccases using bioinformatic approaches has also generated new hybrid enzymes with promising activity (13). Interestingly, the use of bacterial consortia has been proved to be a good alternative for screening of multicopper oxidase activity (14,15).

Despite the increasing interest in bacterial LMCOs, little attention has been paid to plasmid-encoded laccases. We previously published a bioinformatic analysis in which we analyzed the presence of laccase-encoding genes in a variety of bacterial genomes from different species (9). We found that, of the 749 genes identified in finished genomes, 10% were encoded on plasmids; 68% of these genes codify three-domain laccases and 32% two-domain laccases. These findings highlight the importance of plasmids – and other mobile genetic elements (MGE) – as reservoirs of laccases and other enzymes.

In an attempt to identify novel laccases for potential use in industry or bioremediation, we screened a plasmid metagenome obtained from over 50 plasmids purified from bacterial strains carrying high-molecular mass plasmids. This bacterial collection consists of 35 organisms that include both Gram-positive and Gram-negative bacteria belonging to 14 genera, all of them isolated from a biopurification system used for the remediation of pesticide-contaminated waters in Kortrijk, Belgium (16). Exposure of these indigenous bacteria to mixtures of pollutants has fostered their adaptation responses via horizontally acquired MGE. Of the MGE, plasmids are the most abundant and promiscuous ones, representing the main vehicles for horizontal gene transfer via conjugation in bacterial communities in polluted environments (17,18).

Considering the relatively great abundance of laccase-encoding genes in plasmids predicted in silico (9), we hypothesized that the plasmids from the 35 strains are a possible source of laccases. To identify novel LMCOs, we computationally screened this plasmid data set. We were able to identify one full-length gene for a putatively novel LMCO, herein named lac80. The heterologous expression of the codon-optimized version of lac80 in Escherichia coli followed by His-Tag purification allowed us to assess its activity on different substrates, exploring its potential use in bioprocesses.

MATERIALS AND METHODS

Strains and vectors used in this study

Bacterial isolates used in this study were previously obtained from a biopurification system used for pesticide removal from contaminated waters located in Kortrijk, Belgium, operational since 2008 (16). All strains were routinely grown on Luria-Bertani (LB) agar plates or in liquid LB medium (Sigma–Aldrich, Merck, St Louis, MO, USA) at 37 °C and 200 rpm. The biopurification system composition, analysis of pesticide types and concentrations (19), high molecular mass (HMM) plasmid DNA purification, high-throughput sequencing and computational analysis were previously reported (20). Escherichia coli DH5α (Promega, Madison, WI, USA) and E. coli BL21 StarTM (DE3) (Invitrogen Life Technologies, Carlsbad, CA, USA) strains were employed for lac80 cloning and heterologous expression, respectively. They were routinely cultured in LB medium supplemented when appropriate with 100 μg/mL of ampicillin (Sigma–Aldrich, Merck). The pUC59 (GenScript Biotech, Piscataway, NJ, USA) and pET22b(+) (Novagen, Darmstadt, Germany) plasmids were used for lac80 cloning and expression, respectively. The latter plasmid enabled us to introduce a hexahistidine tag (His6-Tag) at the C-terminus of the protein. Molecular biology techniques were performed using the standard protocols (21). All the reagents were purchased from Sigma–Aldrich, Merck unless otherwise stated.

Screening and lac80 sequence analysis

Laccase-like multicopper oxidase (LMCO)-encoding lac80 gene was identified in silico in the present study from the plasmid dataset reported by Martini et al. (20). To retrieve novel LMCOs, profile hidden Markov models (pHMMs) were used as previously described (9). The lac80 gene and aminoacidic sequences were compared to those reported in GenBank (22) and Protein Data Bank (23), respectively. The protein module structure was analyzed using the simple modular architecture research tool (SMART) (24). The presence and location of the signal peptide in lac80 was checked using the neural networks and hidden Markov models trained on Gram-negative and Gram-positive bacteria with SignalP 5.0 (25). Multiple sequence alignment of lac80 with related LMCO sequences was performed using ClustalΩ (26). Laccase gene was also blasted to the Laccase Engineering Database (27) by the basic local alignment search tool (BLASTP) algorithm (28). Phylogenetic relationship was inferred by using the maximum likelihood phylogenetic method. The tree was constructed using MEGA X (29) and a multiple sequence alignment by MUSCLE of Lac80 sequence with a selection of previously characterized and some uncharacterized bacterial LMCOs, and 100 bootstrapping replications were used as a test of phylogeny.

A molecular polymerase chain reaction (PCR)-based method was used to screen the bacterial collection for retrieving this novel LMCO-encoding gene. The strain harbouring lac80 was identified by using two sets of primers (set 1: 5’-CCACCGTCTGGGGTCTTG-3’ and 5’-GTCGATGCGCCGTATTTC-3’, amplicon size 549 bp and set 2: 5’-TCACCGGGCGATGCTGGC-3’ and 5’-GAGGAGGTGATGGCCGAGATC-3’, amplicon size 733 bp). PCR was performed as follows: an initial denaturation step at 94 °C for 4 min, followed by 35 cycles of denaturation at 94 °C for 20 s, annealing at 54 °C for 30 s and extension at 72 °C for 50 s. Then, a final extension at 72 °C for 2 min was used. The sequence of lac80 was submitted to GenBank (22) under the accession number MT130716.

Genetic localization of lac80

In situ lysis gel electrophoresis was performed to obtain further information on the genomic location of lac80 by analyzing the DNA of the bands corresponding to plasmid(s) or chromosome, as previously described (16). The visualized bands were purified using the AccuPrep PCR/gel purification kit (Bioneer, Daejeon, Korea). Then, a semi-nested PCR for lac80 was run to reduce nonspecific amplification of DNA template. Purified samples and primers Lac80F1 (5’-CCACCGTCTGGGGTCTTG-3’) and Lac80R2 (5’-TCACCGGGCGATGCTGGC-3’) were used in the first PCR and Lac80F1 and Lac80R1 (5’-GTCGATGCGCCGTATTTC-3’) in the second PCR run. To exclude the amplification of contaminating DNA, several controls were included in the PCR with DNA recovered from randomly selected positions of the agarose gel.

Codon optimization

The codon optimization strategy employed in the present study is referred to as ‘one amino acid-one codon’. In this method, the most preferred codon of the E. coli expression system for a given amino acid is utilized in the target sequence (30). The sequence of lac80 was obtained from the data set. The OPTIMIZER web server (31) was used for rare codon detection. The E. coli rare codon analyzer2 (32) was utilized for gene sequence optimization. GenScript web server (33) was used to analyze the designed sequence codon adaptation index (CAI).

Cloning and heterologous expression of lac80 in E. coli

The codon-optimized sequence of lac80 with the native signal peptide removed was synthesized by GenScript Biotech (Piscataway, NJ, USA) and cloned into pUC59 plasmid (Leiden, The Netherlands) for DNA amplification. For protein expression, lac80 was cloned from pUC59 into the expression vector pET22b(+) (Novagen Inc, Madison, WI, USA), and introduced into E. coli BL21 StarTM (DE3) cells (Thermo Fisher Scientific, Waltham, MA, USA). Multiple conditions were then evaluated for optimizing Lac80 expression. Briefly, 2 L unbaffled Erlenmeyer flasks containing 750 mL of LB medium (Merck KGaA, Darmstadt, Germany) or richer Terrific broth (TB, in g/L: tryptone 12, yeast extract 24, K2HPO4/L 9.4, KH2PO4 2.2, glycerol 8; components from Merck KGaA) supplemented with ampicillin 100 µg/mL (Merck KGaA) were inoculated with the starter culture (A600 nm~0.1), then incubated at 37 °C and 200 rpm. When the early exponential phase of growth was reached (corresponding to an A600 nm of approx. 0.6 for LB medium and approx. 1 for TB medium), different concentrations of isopropyl-β-d-thiogalactopyranoside (IPTG, from 0.4 to 1 mM) (Merck KGaA) were tested to induce protein expression. At the same time, none or 0.25 mM CuSO4 (Merck KGaA) was added to facilitate metal incorporation into the enzyme active site. Cells were then incubated at 37 or 20 °C under different agitation regimes (200 or 100 rpm or without shaking), and harvested at 0, 2, 4 or 24 h after induction, by centrifuging at 7500×g for 10 min (Thermo Fisher Scientific, Langenselbold, Germany). Multiple combinations of incubation conditions were also evaluated, such as incubation for 4 h after induction at 37 or 20 °C and 200 rpm, followed by overnight incubation at the same temperature but without shaking. Total proteins in the cell-free fermentation broths were concentrated by precipitation with φ(trichloroacetic acid)=10%. Cell pellets were instead resuspended in 20 mM phosphate buffer, pH=6.7, containing 10 µg/mL deoxyribonuclease I (DNaseI) (Merck KGaA), 0.19 mg/mL phenylmethylsulfonylfluoride (PMSF) and 0.7 µg/mL pepstatin (Merck KGaA). Then, cells were disrupted by sonication for 6 cycles of 30 s each on ice, followed by centrifugation at 34 000×g for 60 min at 4 °C (Beckman Coulter Inc., Brea, CA, USA). This enabled us to recover soluble (cytoplasmic) and insoluble (inclusion bodies) fractions, which were then analyzed through: (i) sodium dodecyl sulfate-polyacrylamide (12% m/V) gel electrophoresis (SDS-PAGE) and Coomassie Brilliant Blue staining, (ii) Western blot with anti-His-Tag antibody horseradish peroxidase (HRP) conjugate (Novagen Inc., Madison) and detection by chemiluminescence (ECL Western blotting detection system, GE Healthcare Sciences, Little Chalfont, UK), and (iii) routine laccase activity assay with 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulphonic acid) (ABTS) (Merck KGaA) as substrate (see below). Total protein concentration in the fractions was estimated by the biuret assay (34).

Lac80 purification

Lac80 was purified starting from recombinant E. coli cells grown under the conditions allowing the highest soluble protein production. Hence, glycerol stocks of E. coli BL21 StarTM (DE3)/pET22(b)::lac80 were inoculated into 80 mL LB medium with 100 µg/mL ampicillin, grown overnight at 37 °C and 200 rpm. This pre-culture was then used to inoculate (at A600 nm~0.1) 750 mL of selective LB medium. Flasks were incubated as before until an A600 nm of 0.6 was reached. After induction of protein expression with 0.4 mM isopropyl β-d-1-thiogalactopyranoside (IPTG) and addition of 0.25 mM CuSO4, incubation was prolonged overnight at 20 °C and 100 rpm. Cells were disrupted as reported above and the soluble, cytoplasmic fraction recovered by centrifugation at 34 000×g and 4 °C for 60 min (Beckman Coulter Inc.). C-terminus His6-tagged Lac80 was purified by affinity chromatography on a 5-mL Ni2+-HiTrap chelating affinity column (GE Healthcare Sciences) equilibrated with 20 mM phosphate buffer, pH=6.7, 20 mM imidazole and 300 mM Na2SO4 (Merck KGaA). The recombinant protein was eluted with 20 mM phosphate buffer, pH=6.7, 250 mM imidazole and 300 mM Na2SO4, then loaded onto a size-exclusion PD10 Sephadex G25 column (GE Healthcare Sciences) equilibrated with 20 mM phosphate buffer, pH=6.7. Lac80 was further concentrated with 30 K Amicon Ultra-2 centrifugal filter devices (Merck KGaA).

Protein purity was checked with 10% m/V SDS-PAGE. For molecular mass determination, molecular mass markers (GE Healthcare Sciences) were used. Protein concentration was estimated by densitometric analysis of SDS-PAGE bands using Quantity One analysis software (35) (Bio Rad, Hercules, USA) and spectrophotometrically using Lac80 theoretical molar absorption coefficient at 280 nm calculated based on the amino acid sequence (ε280 nm=58.44 mM-1cm-1).

Enzymatic activity of Lac80

Laccase activity was routinely measured at 25 °C in 1 mL reaction mixture containing 50 mM sodium acetate, pH=5.0, 5 mM ABTS (ε420 nm=36 mM-1cm-1) and 100 μL protein sample. Oxidation of the substrate was monitored with a spectrophotometer (V460; Jasco, Easton, MD, USA) as the change in the absorbance at 420 nm for 5 min. All assays were conducted in triplicate and negative controls without enzyme were run in parallel. One activity unit (U) was defined as the amount of Lac80 required to oxidize 1 μmol of ABTS per minute. The ability of Lac80 to oxidize various phenolic and non-phenolic compounds was checked by the following additional substrates at designed concentrations and wavelengths: 100 mM 2,6-dimethylphenol (DMP) (ε468 nm=49.6 mM-1
cm-1), 100 mM pyrocatechol (ε450 nm=2.21 mM-1cm-1), 100 mM pyrogallol (ε450 nm=4.4 mM-1cm-1), 8 mM vanillic acid (ε316 nm=2.34 mM-1cm-1), 10 mM K4Fe(CN)6 (ε405 nm=0.9 mM-1cm-1), 2 mM tyrosine (ε475 nm=3.6 mM-1cm-1), 5 mM l-3,4-dihydroxyphenylalanine (l-DOPA) (ε475 nm=3.7 mM-1cm-1), 100 mM guaiacol (ε468 nm=12 mM-1cm-1), 10 mM syringic acid (ε300 nm=8.5 mM-1cm-1), 4 mM ferulic acid (ε287 nm=12.4 mM-1cm-1), 1 mM syringaldazine (ε525 nm=65 mM-1cm-1) and 1 mM syringaldehyde (ε320 nm=8.5 mM-1cm-1). The optimal pH for the activity on different substrates was evaluated at 25 °C in a multi-component buffer (10 mM Trizma base, 15 mM sodium carbonate, 15 mM phosphoric acid, 250 mM potassium chloride (36), in the pH range 2.0–8.0. All substrates and buffer components were purchased from Merck KGaA.

RESULTS AND DISCUSSION

Lac80 is a plasmid-encoded three-domain laccase

Polluted sites are hot spots of plasmids potentially carrying catabolic genes (37). To identify novel laccase-like multicopper oxidase (LMCO)-encoding genes in the plasmid dataset from the 35 strains previously isolated from a biopurification system [REMOVED HYPERLINK FIELD] (20), a profile hidden Markov model-based search was performed. One candidate gene, named lac80, was retrieved using this approach. This gene showed 89% identity at nucleotide level with a yet uncharacterized multicopper oxidase from Ochrobactrum anthropi strain OAB (GenBank accession number: CP008819). Lac80 is composed of 502 amino acids, with a predicted signal peptide of 20 amino acids at the N-terminal (Fig. 1a). Predicted isoelectric point and molecular mass of this protein, excluding the signal peptide, were 5.43 and 51.2 kDa, respectively. As reported inTable S1, full-length Lac80 protein showed the highest amino acid sequence identity with two putative multicopper oxidases, one from Alphaproteobacteria (WP_024899901.1, 100% sequence identity) and the other from Ochrobactrum rhizosphaerae (WP_094574672.1, 99.8%). In addition, Lac80 showed 30% sequence identity with the Lac15 from a marine microbial metagenome (Protein Data Bank (PDB): 4F7K_A) and 29.5% identity with a laccase from Pseudomonas thermotolerans (PDB: 6VOW). These two latter proteins were crystallized and characterized as functional LMCOs. The phylogenetic tree shown inFig. 1b confirms the novelty of our enzyme since Lac80 clusters most closely with uncharacterized LMCOs from Paracoccus sp., Ochrobactrum sp. and Rhizobiales. These findings were also confirmed by the best protein matches identified by the Laccase Engineering Database (LccED,https://lcced.biocatnet.de/) (27) by the basic local alignment search tool (BLASTP) algorithm for laccase-like protein.

Fig. 1 In silico analysis of Lac80 protein sequence: a) predicted signal peptide (SP), oxidase type 3 domain (MCO-3) PF07732, oxidase type 1 domain (MCO-1) PF00394 and oxidase type 2 domain (MCO-2) PF07731 are indicated. Residues involved in the coordination of the copper atoms are highlighted in red, and b) maximum likelihood phylogenetic tree of Ochrobactrum sp. BF15 putative laccase-like multicopper oxidase (LMCO) based on reference protein sequences of characterized and some uncharacterized bacterial laccases from Protein Databases (22,23). The percentage of trees in which the associated taxa cluster together is shown next to the branches. MCO=multicopper oxidase, BO=bilirubin oxidase, PO=polyphenol oxidase. Solid black circle indicates the protein characterized in this study
FTB-59-519-f1

In silico sequence analysis of Lac80 identified the three conserved copper-oxidase domains characteristic of laccases, with protein families (Pfam) database accession numbers of PF07732, PF00394 and PF07731 (Fig. 1a). Multiple sequence alignment with bacterial laccases or multicopper oxidase sequences reported inFig. S1 revealed that Lac80 contains metal-binding amino acid residues that are conserved in bacterial laccases.

To identify the strain containing lac80, a PCR-based screening (see Materials and Methods) was performed using gDNA from the 35 strains (38). Of all the tested strains, only one yielded positive PCR results. Identification of the PCR product was corroborated by sequencing. The strain carrying the plasmid-encoded lac80 gene was identified as Ochrobactrum sp. BF15. To determine the genomic localization of lac80, an additional PCR screening was performed using the DNA bands cut from the in situ lysis gel electrophoresis (16) (Fig. 2a andFig. 2b). Positive amplification from plasmid bands (Fig. 2c) and further sequencing confirmed that lac80 is encoded in a plasmid.

Fig. 2 Localization of lac80 in Ochrobactrum sp. BF15: a) in situ gel electrophoresis showing plasmid profile (line 2). The well-known plasmid profile of Sinorhizobium meliloti MVII-1 (line 1) (32) was used to estimate the size of Ochrobactrum sp. BF15 plasmids, b) schematic localization of primers used for lac80 plasmid localization and expected PCR products, and c) agarose gel showing the PCR products obtained with total DNA from Ochrobactrum sp. BF15 or from different plasmids (B1, B2 and B3) purified from the bands. +=positive control, −=negative control, L=ladder
FTB-59-519-f2

Although laccases have already been found in Ochrobactrum spp. (39-41), this is the first report of a plasmid-encoded LMCO in this genus. Ochrobactrum species are aerobic, non-fermenting, Gram-negative bacilli found in a variety of environmental sources such as water, soil, plants and animals. Genome sequencing of Ochrobactrum strains has shown the presence of multiple circular chromosomes and large plasmids (42-46). In addition, plasmid conjugation machinery has been identified in Ochrobactrum plasmids (20,42), and horizontal gene transfer by conjugation was demonstrated in this genus (47,48). Ochrobactrum species was previously found to degrade a wide spectrum of recalcitrant and xenobiotic compounds, such as organophosphates (49), tetrachloroethene (50), dichloro-diphenyl-trichloroethane (DDT) (51), 2,4-dichlorophenoxyacetic acid, endosulfan (52) and aniline (53), among many others. Interestingly, ligninolytic activity was also reported from Ochrobactrum species (39,41). However, this is the first report of a plasmid-encoded functional laccase in the Ochrobactrum genus. Granja-Travez et al. (39) reported the purification and characterization of a laccase, named OcCueO, from Ochrobactrum sp. Although this laccase has the typical three-domain multicopper oxidase structure and displays activity against different substrates, alignment of OcCueO with Lac80 showed a low percentage of identity at the amino acid level (25.75%) (Fig. S2). This is not surprising, considering that OcCueO appears to be encoded in the chromosome, whereas lac80 is located in a plasmid, indicating a possible acquisition via horizontal gene transfer and consequently different evolution pathways.

Heterologous expression and purification of Lac80

The putative LMCO gene was synthesized by optimizing its codon usage for E. coli expression and removing the signal peptide sequence. The synthetic gene was then cloned into the pET22b(+) plasmid, and expressed as a His6-tagged protein in E. coli BL21 StarTM (DE3). Different expression conditions were explored, as detailed in Materials and Methods. Briefly, recombinant cells were grown in two dissimilarly composed cultivation media and increasing concentrations of IPTG were tested to induce Lac80 expression. Following IPTG addition, incubation under diverse temperatures and shaking regimes was compared, together with the addition of CuSO4. Under most of the above conditions, SDS-PAGE and Western blot analyses revealed that His6-Lac80 (molecular mass of 52.5 kDa) accumulated in the inclusion bodies (IBs) in an insoluble form. The recombinant protein was never detected in the extracellular broth, indicating that it could not be secreted. Laccase activity assays that used ABTS as substrate proved that Lac80 accumulated in the IBs was completely inactive, as generally expected for recombinant proteins packed into IBs in E. coli (12,54). This is not surprising, as other LMCOs expressed in E. coli were described to accumulate as inactive protein into IBs. As previously reported, the addition of copper under microaerobic conditions was used to facilitate proper protein folding, a condition that facilitates copper incorporation into the LMCO active sites (55). Thus we grew E. coli BL21 StarTM (DE3)/pET22(b)::lac80 cells in LB medium, supplemented with 0.25 mM CuSO4 at the moment of induction, and incubated overnight at 20 °C and 100 rpm or without shaking. This procedure allowed us to obtain traces of active recombinant Lac80 in the soluble cytoplasmic fraction detectable by both Western blot analysis and activity assay (Fig. 3). Although under these conditions most of the recombinant enzyme was still packed into IBs, approx. 50 µg of soluble Lac80 per L corresponding to approx. 10.6 µg per g cells was produced in the cytoplasmatic fraction. Activity assay (see below) confirmed that this soluble Lac80 was biologically active. Therefore, it was purified by HisTag affinity chromatography, following the procedure described in Materials and Methods. As a precaution, we prepared all the buffers with Na2SO4 rather than with NaCl, since chlorine ions were shown to inhibit laccase activity (8). Approx. 42 µg per L culture and 9 μg per g cells of highly pure Lac80 were recovered from the elution peak, with a purification yield of approx. 84%.

Fig. 3 Heterologous expression of Lac80. Western blot analysis of soluble and insoluble fractions from E. coli BL21 StarTM (DE3)/pET22(b)::lac80 cells was performed. Cells were grown in LB medium, supplemented with 0.25 mM CuSO4 at the same time of IPTG induction, then incubated at 20 °C and 100 rpm and harvested 0, 2, 4 or 24 h after induction. The loaded samples correspond to the soluble or insoluble fractions from cells recovered from V(culture)=1 or 5 mL, respectively. IB=inclusion bodies
FTB-59-519-f3

Lac80 displays activity on different substrates

We tested the ability of Lac80 to oxidize different phenolic and non-phenolic compounds that are typical LMCO substrates. Initial substrate screening was conducted at pH=5.0. Lac80 was active on five of the thirteen screened substrates: the non-phenolic ABTS (activity=1.25 U/mg) and the phenolic 2,6-DMP (activity=10.4 U/mg), pyrocatechol (activity=0.85 U/mg), pyrogallol (activity=3.2 U/mg) and vanillic acid (activity=0.87 U/mg). The highest specific activity at pH=5.0 was reported using 2,6-DMP as substrate, whereas oxidation of the other substrates was overall less efficient. Since for oxidation with LMCOs the optimal pH can vary among substrates, the activity of Lac80 on the five positive substrates was further evaluated at different pH values (from 2.0 to 8.0), as shown inFig. 4. In all cases, the pH-dependent activity showed a bell-shaped curve, with optimum activity in the acidic range. These results show that Lac80 was active in a wide pH range: on 2,6-DMP, for instance, the recombinant LMCO was active in the pH interval from 4.0 to 8.0, whereas on pyrocatechol Lac80 showed activity from pH=3.0 to 8.0. Notably, on this latter substrate, approx. 65 and 53% of activity was also maintained at pH=7.0 and 8.0, respectively. A review of the literature shows that most of the thirteen substrates we tested with Lac80 were indeed oxidized by bacterial laccases in more than half of the reported cases (56,57). Unlike other LMCOs, Lac80 was inactive on tyrosine, thereby excluding the possibility of being a tyrosinase.

Fig. 4 Enzymatic activity of Lac80 under different conditions. Lac80 displayed activity against: a) ABTS, b) DMP, c) pyrocatechol, d) pyrogallol, and e) vanillic acid at 25 °C and in a multi-component buffer in the pH range from 2.0 to 8.0, with 50 μM CuSO4 supplemented to the assay mixture. Values represent mean±standard error of three parallel measurements; when not shown, the error is smaller than the symbol used
FTB-59-519-f4

Interestingly, the recently characterized laccase OcCueO from Ochrobactrum sp. showed activity against ABTS and DMP, two substrates that Lac80 was able to degrade (39). However, OcCueO displayed activity against guaiacol, while Lac80 showed no activity against this substrate. Given the differences at the amino acid level, these results are expectable. It is important to highlight here the prominence of Ochrobactrum species as a source of phenol oxidases.

CONCLUSIONS

Although purification of Lac80 yielded small amounts of active protein, we were able to characterize this laccase using typical substrates. Lac80 shows promising traits such as substrate flexibility and stability at different pH values, making it suitable for industrial purposes. Our further investigations will explore innovative platforms of heterologous expression coupled with using different constructs to overcome the limitations of E. coli system for producing Lac80. In subsequent trials, we will also test other soil-dwelling prokaryotic microbes such as bacilli or streptomycetes, known for their ability to secrete recombinant proteins, to facilitate Lac80 recovery and increase its purification yield. More comprehensive assessment of the biotechnological potential of Lac80 will be possible once the issue of its supply has been solved.

Finally, our study highlights the power of massive sequencing techniques for the discovery of novel oxidizing enzymes in still poorly investigated bacterial genera, especially those encoded in mobile genetic elements. We show that Ochrobactrum species are a promising source of laccase-like enzymes. A still-open question is the physiological function and ecological role of this plasmid-carried laccase-like multicopper oxidase and the way it contributes to the degradation capabilities of soil microbial communities exposed to pollution.

ACKNOWLEDGEMENTS

The authors would like to honour Prof. Gabriel Favelukes and Prof. Oscar Grau who made valuable contributions to the development of the Institute of Biotechnology and Molecular Biology. The authors are grateful to Paula Giménez and Silvana Tongiani for technical assistance. We are grateful to all lab members for helpful discussions. The authors are grateful to Dr. Susan H. Rogers for editing the final version of the manuscript.

Notes

[1] Financial disclosure FUNDING

This work was supported by the National Scientific and Technical Research Council of Argentina (Consejo Nacional de Investigaciones Científicas y Técnicas—CONICET, Argentina) [PIP2015-0700] and the Ministry of Science, Technology and Productive Innovation (Ministerio de Ciencia, Tecnolología e Innovación Productiva—MinCyT, Argentina) [PICT2017-2833 and PICT2017-2371], and in part by the European Commission's 7th Framework Programme (grant number: MetaExplore 222625). MCM was supported by fellowships from CONICET. MP, AL and MFDP are researchers at CONICET. CC was a student of the PhD course in Biotechnologies, Biosciences and Surgical Technologies at University of Insubria, Varese, Italy. LA and IMM were in part supported by Slovenian Research Agency program grant P4-0116 and in part by MetaExplore 222625.

[2] Conflicts of interest CONFLICT OF INTEREST

The authors have no conflicts of interest to declare that are relevant to the content of this article.

SUPPLEMENTARY MATERIAL

REFERENCES

1 

Janusz G, Pawlik A, Świderska-Burek U, Polak J, Sulej J, Jarosz-Wilkołazka A, et al. Laccase properties, physiological functions, and evolution. Int J Mol Sci. 2020;21(3):966. https://doi.org/10.3390/ijms21030966 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32024019

2 

Mate DM, Alcalde M. Laccase: Amulti-purpose biocatalyst at the forefront of biotechnology. Microb Biotechnol. 2017;10(6):1457–67. https://doi.org/10.1111/1751-7915.12422 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27696775

3 

Reiss R, Ihssen J, Richter M, Eichhorn E, Schilling B, Thöny-Meyer L. Laccase versus laccase-like multi-copper oxidase: a comparative study of similar enzymes with diverse substrate spectra. PLoS One. 2013;8(6):e65633. https://doi.org/10.1371/journal.pone.0065633 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/23755261

4 

Liu C, Zhang W, Qu M, Pan K, Zhao X. Heterologous expression of laccase from Lentinula edodes in Pichia pastoris and its application in degrading rape straw. Front Microbiol. 2020;11:1086. https://doi.org/10.3389/fmicb.2020.01086 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32528453

5 

Tülek A, Karataş E, Çakar MM, Aydın D, Yılmazcan Ö, Binay B. Optimisation of the production and bleaching process for a new Laccase from Madurella mycetomatis, expressed in Pichia pastoris: From secretion to yielding prominent. Mol Biotechnol. 2021;63(1):24–39. https://doi.org/10.1007/s12033-020-00281-9 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33058020

6 

Tülek A, Yıldırım D, Aydın D, Binay B. Highly-stable Madurella mycetomatis laccase immobilized in silica-coated ZIF-8 nanocomposites for environmentally friendly cotton bleaching process. Colloids Surf B Biointerfaces. 2021;202:111672. https://doi.org/10.1016/j.colsurfb.2021.111672 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33690061

7 

Yadav D, Ranjan B, Mchunu N, Le Roes-Hill M, Kudanga T. Enhancing the expression of recombinant small laccase in Pichia pastoris by a double promoter system and application in antibiotics degradation. Folia Microbiol (Praha). 2021;66:917–30. https://doi.org/10.1007/s12223-021-00894-w PubMed: http://www.ncbi.nlm.nih.gov/pubmed/34216383

8 

Ausec L, Črnigoj M, Šnajder M, Ulrih NP, Mandic-Mulec I. Characterization of a novel high-pH-tolerant laccase-like multicopper oxidase and its sequence diversity in Thioalkalivibrio sp. Appl Microbiol Biotechnol. 2015;99(23):9987–99. https://doi.org/10.1007/s00253-015-6843-3 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26227413

9 

Ausec L, Zakrzewski M, Goesmann A, Schlüter A, Mandic-Mulec I. Bioinformatic analysis reveals high diversity of bacterial genes for laccase-like enzymes. PLoS One. 2011;6(10):e25724. https://doi.org/10.1371/journal.pone.0025724 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/22022440

10 

Berini F, Verce M, Ausec L, Rosini E, Tonin F, Pollegioni L, et al. Isolation and characterization of a heterologously expressed bacterial laccase from the anaerobe Geobacter metallireducens. Appl Microbiol Biotechnol. 2018;102(5):2425–39. https://doi.org/10.1007/s00253-018-8785-z PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29380032

11 

Casciello C, Tonin F, Berini F, Fasoli E, Marinelli F, Pollegioni L, et al. A valuable peroxidase activity from the novel species Nonomuraea gerenzanensis growing on alkali lignin. Biotechnol Rep (Amst). 2017;13:49–57. https://doi.org/10.1016/j.btre.2016.12.005 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28352563

12 

Ausec L, Berini F, Casciello C, Cretoiu MS, van Elsas JD, Marinelli F, et al. The first acidobacterial laccase-like multicopper oxidase revealed by metagenomics shows high salt and thermo-tolerance. Appl Microbiol Biotechnol. 2017;101(15):6261–76. https://doi.org/10.1007/s00253-017-8345-y PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28589226

13 

Itoh N, Hayashi Y, Honda S, Yamamoto Y, Tanaka D, Toda H. Construction and characterization of a functional chimeric laccase from metagenomes suitable as a biocatalyst. AMB Express. 2021;11(1):90. https://doi.org/10.1186/s13568-021-01248-y PubMed: http://www.ncbi.nlm.nih.gov/pubmed/34146179

14 

Dai X, Lv J, Yan G, Chen C, Guo S, Fu P. Bioremediation of intertidal zones polluted by heavy oil spilling using immobilized laccase-bacteria consortium. Bioresour Technol. 2020;309:123305. https://doi.org/10.1016/j.biortech.2020.123305 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/32325376

15 

Guo G, Liu C, Hao J, Tian F, Ding K, Zhang C, et al. Development and characterization of a halo-thermophilic bacterial consortium for decolorization of azo dye. Chemosphere. 2021;272:129916. https://doi.org/10.1016/j.chemosphere.2021.129916 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/33601203

16 

Martini MC, Albicoro FJ, Nour E, Schlüter A, van Elsas JD, Springael D, et al. Characterization of a collection of plasmid-containing bacteria isolated from an on-farm biopurification system used for pesticide removal. Plasmid. 2015;80:16–23. https://doi.org/10.1016/j.plasmid.2015.05.001 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25957823

17 

Garcillán-Barcia MP, Francia MV, de la Cruz F. The diversity of conjugative relaxases and its application in plasmid classification. FEMS Microbiol Rev. 2009;33(3):657–87. https://doi.org/10.1111/j.1574-6976.2009.00168.x PubMed: http://www.ncbi.nlm.nih.gov/pubmed/19396961

18 

Harrison E, Brockhurst MA. Plasmid-mediated horizontal gene transfer is a coevolutionary process. Trends Microbiol. 2012;20(6):262–7. https://doi.org/10.1016/j.tim.2012.04.003 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/22564249

19 

Dealtry S, Holmsgaard PN, Dunon V, Jechalke S, Ding GC, Krögerrecklenfort E, et al. Shifts in abundance and diversity of mobile genetic elements after the introduction of diverse pesticides into an on-farm biopurification system over the course of a year. Appl Environ Microbiol. 2014;80(13):4012–20. https://doi.org/10.1128/AEM.04016-13 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24771027

20 

Martini MC, Wibberg D, Lozano M, Torres Tejerizo G, Albicoro FJ, Jaenicke S, et al. Genomics of high molecular weight plasmids isolated from an on-farm biopurification system. Sci Rep. 2016;6:28284. https://doi.org/10.1038/srep28284 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/27321040

21 

Sambrook J, Russel DW. Molecular cloning: A laboratory manual, vol 2. Cold Spring Harbor, NY, USA: Cold Spring Harbor Laboratory Press; 2001.

22 

Clark K, Karsch-Mizrachi I, Lipman DJ, Ostell J, Sayers EW. GenBank. Nucleic Acids Res. 2016;44 D1:D67–72. https://doi.org/10.1093/nar/gkv1276 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26590407

23 

Ge H, Xiao Y. Crystal structure of Lac15 from a marine microbial metagenome. Available from:https://www.rcsb.org/structure/4f7k. https://doi.org/

24 

Schultz J, Milpetz F, Bork P, Ponting CP. SMART, a simple modular architecture research tool: Identification of signaling domains. Proc Natl Acad Sci USA. 1998;95(11):5857–64. https://doi.org/10.1073/pnas.95.11.5857 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/9600884

25 

Nielsen H, Engelbrecht J, Brunak S, von Heijne G. A neural network method for identification of prokaryotic and eukaryotic signal peptides and prediction of their cleavage sites. Int J Neural Syst. 1997;8(5-6):581–99. https://doi.org/10.1142/S0129065797000537 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/10065837

26 

Sievers F, Higgins DG. Clustal Omega, accurate alignment of very large numbers of sequences. In: Russell D, editor. Multiple sequence alignment methods: Methods in molecular biology (Methods and protocols), vol. 1079. Totowa, NJ, USA: Humana Press; 2014. pp. 105-16. https://doi.org/10.1007/978-1-62703-646-7_6 https://doi.org/10.1007/978-1-62703-646-7_6

27 

Sirim D, Wagner F, Wang L, Schmid RD, Pleiss J. The Laccase Engineering Database: A classification and analysis system for laccases and related multicopper oxidases. Database (Oxford). 2011;2011:bar006. https://doi.org/10.1093/database/bar006 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21498547

28 

BlastP – Standard Protein BLAST. Bethesda, MD, USA: National Center for Biotecnology Information; 2021. Available from:https://blast.ncbi.nlm.nih.gov/Blast.cgi?PAGE=Proteins.

29 

Kumar S, Stecher G, Li M, Knyaz C, Tamura K. MEGA X: Molecular Evolutionary Genetics Analysis across computing platforms. Mol Biol Evol. 2018;35(6):1547–9. https://doi.org/10.1093/molbev/msy096 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29722887

30 

Zylicz-Stachula A, Zolnierkiewicz O, Sliwinska K, Jezewska-Frackowiak J, Skowron PM. Modified ‘one amino acid-one codon’ engineering of high GC content TaqII-coding gene from thermophilic Thermus aquaticus results in radical expression increase. Microb Cell Fact. 2014;13:7. https://doi.org/10.1186/1475-2859-13-7 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24410856

31 

Puigbò P, Guzmán E, Romeu A, Garcia-Vallvé S. OPTIMIZER: A web server for optimizing the codon usage of DNA sequences. Nucl Acids Res. 2007;35(Web Server issue):W126-31. https://doi.org/10.1093/nar/gkm219 https://doi.org/10.1093/nar/gkm219

32 

Maduro M. E. coli codon usage analyzer 2.1. Riverside, CA; USA: University of California; 2003. Available from:http://www.faculty.ucr.edu/~mmaduro/codonusage/usage.htm.

33 

GenScript Web Server. Piscataway, NJ; USA; 2021. Available from:http://www.genscript.com.

34 

Gornall AG, Bardawill CJ, David MM. Determination of serum proteins by means of the biuret reaction. J Biol Chem. 1949;177(2):751–66. https://doi.org/10.1016/S0021-9258(18)57021-6 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/18110453

35 

Quantity One software, Bio-Rad Laboratories, Inc., Hercules, CA, USA; 2021. Available from:https://www.bio-rad.com/it-it/product/quantity-one-1-d-analysis-software?ID=1de9eb3a-1eb5-4edb-82d2-68b91bf360fb.

36 

Harris CM, Pollegioni L, Ghisla S. pH and kinetic isotope effects in d-amino acid oxidase catalysis. Eur J Biochem. 2001;268(21):5504–20. https://doi.org/10.1046/j.1432-1033.2001.02462.x PubMed: http://www.ncbi.nlm.nih.gov/pubmed/11683874

37 

Dealtry S, Ding GC, Weichelt V, Dunon V, Schlüter A, Martini MC, et al. Cultivation-independent screening revealed hot spots of IncP-1, IncP-7 and IncP-9 plasmid occurrence in different environmental habitats. PLoS One. 2014;9(2):e89922. https://doi.org/10.1371/journal.pone.0089922 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/24587126

38 

Martini MC, Quiroga MP, Pistorio M, Lagares A, Centrón D, Del Papa MF. Novel environmental class 1 integrons and cassette arrays recovered from an on-farm bio-purification plant. FEMS Microbiol Ecol. 2018;94(3):fix190. https://doi.org/10.1093/femsec/fix190 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29300936

39 

Granja-Travez RS, Wilkinson RC, Persinoti GF, Squina FM, Fülöp V, Bugg TDH. Structural and functional characterisation of multi-copper oxidase CueO from lignin-degrading bacterium Ochrobactrum sp. reveal its activity towards lignin model compounds and lignosulfonate. FEBS J. 2018;285(9):1684–700. https://doi.org/10.1111/febs.14437 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29575798

40 

Li Y, Zuo W, Li Y, Wang X. Cloning of multicopper oxidase gene from Ochrobactrum sp. 531 and characterization of its alkaline laccase activity towards phenolic substrates. Adv Biol Chem. 2021;2(3):248–55. https://doi.org/10.4236/abc.2012.23031

41 

Tsegaye B, Balomajumder C, Roy P. Isolation and characterization of novel lignolytic, cellulolytic, and hemicellulolytic bacteria from wood-feeding termite Cryptotermes brevis. Int Microbiol. 2019;22(1):29–39. https://doi.org/10.1007/s10123-018-0024-z PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30810928

42 

Chain PS, Lang DM, Comerci DJ, Malfatti SA, Vergez LM, Shin M, et al. Genome of Ochrobactrum anthropi ATCC 49188 T, a versatile opportunistic pathogen and symbiont of several eukaryotic hosts. J Bacteriol. 2011;193(16):4274–5. https://doi.org/10.1128/JB.05335-11 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/21685287

43 

Poszytek K, Karczewska-Golec J, Ciok A, Decewicz P, Dziurzynski M, Gorecki A, et al. Genome-guided characterization of Ochrobactrum sp. POC9 enhancing sewage sludge utilization—Biotechnological potential and biosafety considerations. Int J Environ Res Public Health. 2018;15(7):1501. https://doi.org/10.3390/ijerph15071501 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/30013002

44 

Tobias NJ, Mishra B, Gupta DK, Ke LP, Thines M, Bode HB. Draft genome sequence of Ochrobactrum anthropi strain ML7 isolated from soil samples in Vinhphuc Province, Vietnam. Genome Announc. 2015;3(2):e00218–15. https://doi.org/10.1128/genomeA.00218-15 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25838481

45 

Trujillo ME, Willems A, Abril A, Planchuelo AM, Rivas R, Ludeña D, et al. Nodulation of Lupinus albus by strains of Ochrobactrum lupini sp. nov. Appl Environ Microbiol. 2005;71(3):1318–27. https://doi.org/10.1128/AEM.71.3.1318-1327.2005 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/15746334

46 

Wang X, Jin D, Zhou L, Zhang Z. Draft genome sequence of Ochrobactrum anthropi strain W13P3, a halotolerant polycyclic aromatic hydrocarbon-degrading bacterium. Genome Announc. 2015;3(4):e00867–15. https://doi.org/10.1128/genomeA.00867-15 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26227610

47 

de la Cruz-Perera CI, Ren D, Blanchet M, Dendooven L, Marsch R, Sørensen SJ, et al. The ability of soil bacteria to receive the conjugative IncP1 plasmid, pKJK10, is different in a mixed community compared to single strains. FEMS Microbiol Lett. 2013;338(1):95–100. https://doi.org/10.1111/1574-6968.12036 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/23106414

48 

Soda S, Otsuki H, Inoue D, Tsutsui H, Sei K, Ike M. Transfer of antibiotic multiresistant plasmid RP4 from Escherichia coli to activated sludge bacteria. J Biosci Bioeng. 2008;106(3):292–6. https://doi.org/10.1263/jbb.106.292 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/18930008

49 

Talwar MP, Mulla SI, Ninnekar HZ. Biodegradation of organophosphate pesticide quinalphos by Ochrobactrum sp. strain HZM. J Appl Microbiol. 2014;117(5):1283–92. https://doi.org/10.1111/jam.12627 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25155583

50 

Tabernacka A, Zborowska E, Pogoda K, Żołądek M. Removal of tetrachloroethene from polluted air by activated sludge. Environ Technol. 2019;40(4):470–9. https://doi.org/10.1080/09593330.2017.1397759 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/29098945

51 

Pan X, Xu T, Xu H, Fang H, Yu Y. Characterization and genome functional analysis of the DDT-degrading bacterium Ochrobactrum sp. DDT-2. Sci Total Environ. 2017;592:593–9. https://doi.org/10.1016/j.scitotenv.2017.03.052 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28320527

52 

Seralathan MV, Sivanesan S, Nargunanathan S, Bafana A, Kannan K, Chakrabarti T. Chemotaxis-based endosulfan biotransformation: Enrichment and isolation of endosulfan-degrading bacteria. Environ Technol. 2015;36(1):60–7. https://doi.org/10.1080/09593330.2014.937464 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/25409584

53 

Yang L, Ying C, Fang N, Zhong Y, Zhao-Xiang Z, Yun S. Identification and characterization of a high efficiency aniline resistance and degrading bacterium MC-01. Appl Biochem Biotechnol. 2017;182(1):41–54. https://doi.org/10.1007/s12010-016-2309-z PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28050740

54 

Berini F, Presti I, Beltrametti F, Pedroli M, Vårum KM, Pollegioni L, et al. Production and characterization of a novel antifungal chitinase identified by functional screening of a suppressive-soil metagenome. Microb Cell Fact. 2017;16(1):16. https://doi.org/10.1186/s12934-017-0634-8 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/28137256

55 

Durão P, Chen Z, Fernandes AT, Hildebrandt P, Murgida DH, Todorovic S, et al. Copper incorporation into recombinant CotA laccase from Bacillus subtilis: Characterization of fully copper loaded enzymes. J Biol Inorg Chem. 2008;13(2):183–93. https://doi.org/10.1007/s00775-007-0312-0 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/17957391

56 

Ihssen J, Reiss R, Luchsinger R, Thöny-Meyer L, Richter M. Biochemical properties and yields of diverse bacterial laccase-like multicopper oxidases expressed in Escherichia coli. Sci Rep. 2015;5:10465. https://doi.org/10.1038/srep10465 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/26068013

57 

Reiss R, Ihssen J, Richter M, Eichhorn E, Schilling B, Thöny-Meyer L. Laccase versus laccase-like multi-copper oxidase: A comparative study of similar enzymes with diverse substrate spectra. PLoS One. 2013;8(6):e65633. https://doi.org/10.1371/journal.pone.0065633 PubMed: http://www.ncbi.nlm.nih.gov/pubmed/23755261

Floating objects

Table S1 BLASTP alignment of the putative laccase-like multicopper oxidase protein sequence (Lac80) encoded by a plasmid-carried gene from Ochrobactrum sp. BF15. The ten best protein matches identified in NCBI by BLASTP (28) algorithm for Lac80 protein are shown
AccessionDescriptionProtein lengthOrganismIdentity/%E valueMax scoreTotal scoreQuery cover/%
WP_024899901.1multicopper oxidase domain-containing protein502Alphaproteobacteria100.00010001000100.00
WP_094574672.1multicopper oxidase domain-containing protein502Ochrobactrum rhizosphaerae99.800999999100.00
WP_138787545.1multicopper oxidase domain-containing protein502Ochrobactrum haematophilum99.800999999100.00
WP_036709831.1multicopper oxidase domain-containing protein502Paracoccus sanguinis99.600997997100.00
WP_155044758.1multicopper oxidase domain-containing protein502unclassified Paracoccus98.800993993100.00
WP_085378474.1multicopper oxidase domain-containing protein493Paracoccus contaminans91.830877877100.00
WP_126156115.1multicopper oxidase domain-containing protein493Paracoccus haematequi91.630873873100.00
WP_121583054.1multicopper oxidase domain-containing protein499Mesorhizobium sp. YM1C-6-290.040870870100.00
WP_114350933.1multicopper oxidase domain-containing protein493Paracoccus lutimaris91.430868868100.00
WP_122520894.1multicopper oxidase domain-containing protein499Pannonibacter phragmitetus90.040867867100.00
Fig. S1 Partial amino acid multiple sequence alignment, showing 10 bacterial laccase-like multicopper oxidases. Metal-binding residues are coloured in blue; other conserved residues are highlighted in red, whereas mismatches are in yellow. Numbering is based on Lac80 sequence as reference. Protein data Bank (23) or GenBank (22) accession codes are as follows: 1=QQZ00696.1, 2=6EVG|A, 3=P07788, 4=P36649, 5=A0A0E0XT94, 6=(GB) WP_011709064.1, 7=Q88C03, 8=(GB) WP_013012601.1, 9=P12374, 10=E1ACR6. BO=bilirubin oxidase, MMM=marine microbial metagenome.
FTB-59-519-fS.jpg
Fig. S2 Sequence alignment of Lac80 with OcCueO from Ochrobactrum sp. Alignment was carried out using Clustal Omega online tool (26). The percentage of identity is 25.5%. Metal-binding residues are coloured in blue; other conserved residues among the bacterial laccase-like multicopper oxidases shown inFig. S1 are highlighted in red, whereas mismatches are in yellow. Uniprot (23) or GenBank (22) accession codes as follows: Lac80 (QQZ00696.1), OcCueO (6EVG|A)
FTB-59-519-fS.jpg

This display is generated from NISO JATS XML with jats-html.xsl. The XSLT engine is libxslt.