Next Article in Journal
An Engineered Protein-Based Building Block (Albumin Methacryloyl) for Fabrication of a 3D In Vitro Cryogel Model
Next Article in Special Issue
Study on Screening Criteria of Gel-Assisted Polymer and Surfactant Binary Combination Flooding after Water Flooding in Strong Edge Water Reservoirs: A Case of Jidong Oilfield
Previous Article in Journal
Preparation, Characterization of Pregabalin and Withania coagulans Extract-Loaded Topical Gel and Their Comparative Effect on Burn Injury
Previous Article in Special Issue
Types and Performances of Polymer Gels for Oil-Gas Drilling and Production: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gene Expression and Characterization of Iturin A Lipopeptide Biosurfactant from Bacillus aryabhattai for Enhanced Oil Recovery

1
Department of Biotechnology, KLE Technological University, Hubballi 580031, India
2
Department of Chemistry, Indian Institute of Technology, Dharwad 580011, India
3
Department of Biosciences and Bioengineering, Indian Institute of Technology, Dharwad 580011, India
4
Department of Mechanical Engineering, College of Engineering, King Khalid University, Abha 61421, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Submission received: 25 May 2022 / Revised: 16 June 2022 / Accepted: 20 June 2022 / Published: 25 June 2022
(This article belongs to the Special Issue Gels for Oil Drilling and Enhanced Recovery)

Abstract

:
Biosurfactants are eco-friendly surface-active molecules recommended for enhanced oil recovery techniques. In the present study, a potential lipopeptide (biosurfactant) encoding the iturin A gene was synthesized from Bacillus aryabhattai. To improvise the yield of the lipopeptide for specific applications, current research tends toward engineering and expressing recombinant peptides. An iturin A gene sequence was codon-optimized, amplified with gene-specific primers, and ligated into the pET-32A expression vector to achieve high-level protein expression. The plasmid construct was transformed into an E. coli BL21 DE3 host to evaluate the expression. The highly expressed recombinant iturin A lipopeptide was purified on a nickel nitrilotriacetic acid (Ni-NTA) agarose column. Sodium dodecyl sulfate polyacrylamide gel electrophoresis (SDS-PAGE) revealed that the purity and molecular mass of iturin A was 41 kDa. The yield of recombinant iturin A was found to be 60 g/L with a 6.7-fold increase in comparison with our previously published study on the wild strain. The approach of cloning a functional fragment of partial iturin A resulted in the increased production of the lipopeptide. When motor oil was used, recombinant protein iturin A revealed a biosurfactant property with a 74 ± 1.9% emulsification index (E24). Purified recombinant protein iturin A was characterized by mass spectrometry. MALDI-TOF spectra of trypsin digestion (protein/trypsin of 50:1 and 25:1) showed desired digested mass peaks for the protein, further confirming the identity of iturin A. The iturin A structure was elucidated based on distinctive spectral bands in Raman spectra, which revealed the presence of a peptide backbone and lipid. Recombinant iturin A was employed for enhanced oil recovery through a sand-packed column that yielded 61.18 ± 0.85% additional oil. Hence, the novel approach of the high-level expression of iturin A (lipopeptide) as a promising biosurfactant employed for oil recovery from Bacillus aryabhattai is not much reported. Thus, recombinant iturin A demonstrated its promising ability for efficient oil recovery, finding specific applications in petroleum industries.

Graphical Abstract

1. Introduction

Biosurfactants are produced as extracellular molecules on microbial cell surfaces. These are surface-active molecules that reduce the surface tension between hydrophobic and hydrophilic regions and are produced by several microbes, such as bacteria, fungi, and yeast [1]. Biosurfactants are biochemically made up of hydrophilic segments, such as carbohydrates, polar peptides, or acids, and hydrophobic segments, usually lipids [2]. A lipopeptide is a specific class of biosurfactant that has a lipid moiety attached to a peptide and is produced as part of microbial metabolism. Lipopeptides have a wide range of applications, especially those produced by Bacillus species, which have been shown to exhibit antibacterial and antifungal activities [3]. Lipopeptides produced by bacterial species can be classified into three types, namely, surfactin, iturin, and fengycin. Out of these three lipopeptides, iturin is comparatively smaller and is amphiphilic in nature. It has a peptide linked to the fatty acid chain. As lipopeptides are molecules of biological origin, they are biodegradable, nontoxic, and nonpolluting biomolecules; hence. they are environmentally acceptable [4].
Iturin A is a type of cyclic lipopeptide produced by Bacillus species as a secondary molecule and acts as an antibiotic [5]. Iturin A has been known to show inhibitory activity by limiting the growth of fungi [6]. The lipopeptide nature of the biosurfactant has shown thermal, pH, and salinity stability [7]. Many researchers are interested in replacing synthetic surfactants with biosurfactants; however, a genuine problem still exists in the economics of production, lower yield, and costly downstream and recovery process of a biosurfactant [8]. However, with the application of statistical design methods, such as response surface methodology (RSM), the yields of biosurfactant production has improved by 2.51-fold using crude oil [7]. To reduce the cost involved in the production of biosurfactants, it is essential to construct hyperproducing or mutant strains with better yields [8]. Improved yield combined with a cost-effective carbon source can result in lesser production cost [9].
The development of recombinant biosurfactants is gaining scientific attention in recent times due to industrial applications. A strategy for improving the gene expression that encodes a biosurfactant is receiving focused interest to improve the yields and modify the chemical attributes [10]. The mycosubtilin and iturin A operons have been determined as operons of the iturin lipopeptide family. The lpa-14 gene is a homologue of the sfp gene, which is responsible for the biosynthesis of surfactin and iturin A in B. subtilis RB14 [11]. The Escherichia coli lac operator and the promoter core region of the penicillinase gene of B. licheniformis present in the hybrid promoter P spac was reported to increase lipopeptide biosynthesis fivefold [12]. Biosurfactants are amphipathic molecules that enable them to perform a vital function. Recently, it has been proved that biosurfactants can solubilize and mobilize organic compounds adsorbed on soil constituents. Biosurfactants have shown good results for the decontamination of soil, as it has unique properties [13]. Biosurfactants can also be employed in biomedical applications owing to their antimicrobial and antiviral activities [14]. The lipopeptide surfactin is one of the most well-known and studied surfactants [15]. A lipopeptide can be used for antimycoplasmal, antibacterial, and antiviral activities. It can also be used for antiadhesive and anti-inflammatory applications. It has also shown good results for environmental applications, such as oil recovery, in environmental bioremediation, and it also accelerates the biodegradation of hydrocarbons [16]. With the help of genome shuffling, 179.22 mg/L of iturin A was obtained from Bacillus amyloliquefaciens, which had a 2.03-fold increase compared with a wild strain [17]. Iturin A production was 330 µg/mL from the recombinant Bacillus subtilis strain as compared with 110 µg/mL from a wild strain, which indicated a threefold increase [6,17]. A biosurfactant produced from Bacillus species was characterized previously by mass spectrometry and Raman spectroscopy [17,18].
However, not many reports are available on the high-level expression of iturin A (lipopeptide) as a promising biosurfactant from Bacillus aryabhattai. The current study reports the design and synthesis of the lpa-14 gene responsible for iturin A production and cloning in the pET-32A vector, followed by the overexpression in E. coli BL21 DE3. The structural characterization of a purified recombinant iturin A protein was carried out by Raman spectroscopy, and molecular mass was analyzed by mass spectroscopy (MALDI-TOF MS). The biosurfactant activity of expressed iturin A was demonstrated in a sand-packed column for enhanced oil recovery.

2. Materials and Methods

2.1. Chemicals, Vectors, and Bacterial Strains

All the chemicals and nickel nitrilotriacetic acid (Ni-NTA) resin column used in the present research were purchased from Sigma-Aldrich Pvt Ltd. (Burlington, MA, USA) and Merck and Co. Inc. (Rahway, NJ, USA). The iturin A gene was synthesized from Bacillus aryabhattai. The pET-32A vector, pUC19 vector, E. coli BL21 DE3, and E. coli DH5α were used for cloning, sequencing, and expression activities.

2.2. Genome Sequence Identification

Iturin A production by B. aryabhattai was identified based on the results obtained from antiSMASH, which predicts the genome clusters producing secondary metabolites [18]. Iturin A synthesis is assisted by the enzyme phosphopantetheinyl transferase coded by the lpa-14 gene. The FASTA sequence of the lpa-14 gene was retrieved from the National Center Biotechnology Information (NCBI).

Gene Synthesis, Amplification, and Cloning

From the synthesized codon-optimized DNA template, the iturin A gene was amplified using the gene-specific primers and ligated into the pET-32A expression vector. The ligated product was transformed into an E. coli DH5α host. Using BamHI and EcoRI, restriction digestion was carried out for the confirmation of the gene insert. The FASTA sequence of the iturin A gene was retrieved from the NCBI (Accession no. QHD44466.1) database, and then it was reverse-translated using the Transeq tool, European Molecular Biology Laboratory (EMBL) [19]. The translated protein sequences were further analyzed for the presence of GC content and subjected to codon optimization for E. coli expression by interchanging certain nucleotide sequences without altering the protein sequence for the enhanced expression of the protein. The gene was amplified using synthesized T7 gene-specific primers (forward primer 5′-ATGAAAATTTACGGAGTATA-3′ and reverse primer 5′-TTATAACAGCTCTTCATACGTT-3′) from the synthesized DNA template. The amplified gene was cloned into the pUC19 vector.
Upon restriction digestion, the cloned insert was confirmed through sequencing. Further, the restricted insert was subcloned into the pET-32A bacterial expression vector; for this activity, BamHI and EcoRI with GGATCC and GAATTC restriction sites were employed. The pET-32A vector used for cloning was designed for the high-level expression of recombinant peptide sequences and was fused with a 109 amino acid thioredoxin protein tag, with cleavable His-tag and S-tag. The plasmid construct with a concentration of 106 ng/µL was transformed into E. coli BL21 DE3 chemical competent cells using heat shock to evaluate the expression. The transformants were subjected to lysate/colony PCR using gene-specific primers to confirm the presence of the insert (iturin A gene). The positive transformants were grown in Luria–Bertani (LB) media and evaluated for protein expression.

2.3. Expression of the Iturin A Lipopeptide

The pET-32A-vector-transformed fresh colony was selected and cultured on an LB agar plate. The starter culture was prepared by inoculating the culture into an LB medium containing ampicillin at 100 µg/mL. The medium was incubated at 37 ± 2 °C under shaking conditions (160 rpm) for 16 h. An amount of 10 mL of starter culture was added to the flask containing a 1 L LB medium with ampicillin and incubated at 37 ± 2 °C. The flask was kept under shaking conditions (200 rpm) to achieve optical density (OD) between 0.55 and 0.7 at 600 nm. Further, the flask was cooled to 16 °C, and isopropyl β-D-1-thiogalactopyranoside (IPTG) at 1 mM was added. The flask was kept at 15 °C at 180 rpm for 16–18 h. The induced cell culture was centrifuged at 4000 rpm at 4 °C for 20 min, and the pellet was collected and stored at −80 °C until further use. For protein purification, the cell pellet was resuspended in ice-cold lysis buffer (20 mM Na2HPO4, 500 mM NaCl, 10 mM imidazole, 2.5% v/v glycerol, and 0.1% v/v Triton), and to this, 80 mg of lysozyme, 20 µL of benzamidine, and 20 µL of phenylmethylsulfonyl fluoride (PMSF) were added. The solution was subjected to sonication for 1 min at 60 amplitude, and this cycle was repeated six times. The solution was centrifuged at 4900 rpm for 60 min at 4 °C. The obtained supernatant was collected in 2 mL tubes and centrifuged again at 4 °C at 11,000 rpm for 60 min to remove any precipitate or suspended material. The clear supernatant was further subjected to purification.

2.4. Purification of the Recombinant Iturin A Lipopeptide

For purification, the obtained supernatant containing protein was loaded onto an agarose column containing nickel nitrilotriacetic acid (Ni-NTA) resin pretreated with lysis buffer for purification of the His-tagged iturin A recombinant protein. The column was washed with 30 mL of wash buffer (20 mM Na2HPO4, 500 mM NaCl, 30 mM imidazole, 2.5% v/v glycerol, pH 7.40) to remove any nontagged proteins. An elution buffer (20 mM Na2HPO4, 300 mM NaCl, 300 mM imidazole, 2.5% v/v glycerol, pH 7.40) was used to elute the His-tagged recombinant protein from the column. The purity of the protein was assessed by sodium dodecyl sulfate polyacrylamide gel (12.5%) electrophoresis (SDS-PAGE) [8]. The expressed iturin A concentration was determined by the Bradford assay (using BSA as the standard protein). The absorbance measurement was recorded at A280 nm (molar extinction coefficient was obtained from the Expasy ProtParam tool using the primary sequence of iturin A; http://web.expasy.org/protparam/, accessed on 15 October 2021).

2.5. Evaluation of Biosurfactant Activity

Iturin A was examined for activities related to biosurfactants by various assays. The oil spread assay was carried out according to the method of Yaraguppi et al. (2020) [7]. An amount of 40 mL of distilled water was taken in a Petri plate, and to this, 25 µL of motor oil was added. An amount of 10 µL of purified recombinant iturin A was added to the oil surface, and oil displacement was measured. The dispersion index is calculated by measuring the diameter of oil after dispersion, and the oil dispersion area was calculated by using the equation = π r 2 , where r is the radius. The biosurfactant activity of iturin A was also assessed by measuring the emulsification index (E24) [7]. Briefly, 3 mL of a recombinant protein (cell-free supernatant) was added to 3 mL motor oil, mixed rigorously (10 min) and allowed to stand for 24 h. The height of the emulsion layer (mm) divided by the total height of the solution (mm) multiplied by 100 indicated the percentage emulsification index [20].

2.6. Analytical Characterization

2.6.1. Mass Spectrometry Analysis

To confirm the identity of the recombinant protein, mass spectrometric (MS) analysis was performed on a tryptic digest of iturin A. The digestion of protein was carried out by using Trypsin-ultra™ Mass Spectrometry Grade from New England Biolabs (NEB, Hertfordshire, UK). Two trypsin digestion reactions for the protein were set up with a protein/trypsin ratio of 50:1 and 25:1 each containing 1 µL of trypsin (1 µg/µL), 2.5 µL of 2X trypsin reaction buffer, 2.5 µL of ultrapurified water, and 50 and 25 µL of protein (1 µg/µL), respectively. These reactions were incubated under shaking conditions overnight (16 h) at 37 ± 2 °C and stored at −80 °C. Samples for Matrix Assisted Laser Desorption/Ionization Time-of-Flight (MALDI-TOF) MS analysis were prepared by desalting the reaction mixtures with C-18 Zip-Tip (Millipore, CA, USA). MALDI-TOF MS was performed at the Indian Institute of Science (IISc, Bangalore, India) Mass Spectrometry Facility, Department of Biological Sciences, Bangalore, using sinapic acid as the matrix.

2.6.2. Raman Spectroscopy Analysis

A Raman alpha 300R (WITec, GmbH, Remscheid, Germany) confocal Raman microscope was used for acquiring spectra of purified protein sample. The equipment consisted of an excitation laser of 532 nm, 50X objective (Zeiss, NA = 0.8), 300 mm spectrograph equipped with 1200 gr/mm grating, and Thermo cooled charge-coupled detector (CCD). Raman-scattered photons were collected through a 50 µm fiber, and the spectra were obtained over the fingerprint region (600–1800 cm−1). The specimen was placed on a quartz window, and a total of 10 spectra were obtained at different locations. Each spectrum was acquired for 5 s and averaged over 10 accumulations. The highest peak normalization was performed using MATLAB (version R2021a). The mean spectrum was computed by averaging the intensities at the Y-axis and keeping the X-axis constant.

2.7. Sand-Packed Column Assay for Enhanced Oil Recovery

To evaluate the biosurfactant nature of recombinant iturin A, sand-packed column experiments were designed for improved oil recovery. Columns were made using 50 mL syringe tubes and were filled with 35 g of sand totally dried at 100 °C in a hot-air oven. Positive and negative controls were performed using SDS solution (1 mg/mL) and ultrapure water, respectively. Each column was flooded with 30 mL of 5% brine solution. The pore volume (PV) of the sand-packed column was calculated by determining the volume of the brine required to saturate the column. PV divided by the volume of the column indicates the porosity (%) of the sand-packed column. Crude oil was flooded into the sand-packed column till it replaced the whole volume of the brine. Hence, the oil in place (OOIP) indicates the volume of crude oil retained in the column [21].
Soi (%) = (OOIP/PV) × 100
Swi (%) = ((PV − OOIP)/PV) × 100
Residual oil saturation was obtained by flooding the brine into the column. Using the following equation, the oil recovered after flooding with the brine and the residual oil saturation was calculated.
Sor (%) = ((OOIP − Sorwf)/OOIP) × 100
The remaining oil was flooded with recombinant iturin A and incubated at 50 ± 2 °C for 24 h. Recovered oil was used to identify the additional oil recovery.
AOR (%) = (Sorif/(OOIP − Sorwf)) × 100
Where Sorwf in mL is residual oil saturation after water flooding, Sorif in mL is oil collected over residual oil saturation after iturin A flooding, Soi % is initial oil saturation, Swi in % is initial water saturation, Sor in % residual oil saturation and AOR in % represents additional oil recovery [22]. AOR % represents the enhanced oil recovery from the sand-packed column.

3. Results and Discussion

3.1. Genome Sequence Identification

The gene track of the lpa-14 nucleotide sequence was identified and retrieved from NCBI and showed the product obtained from the gene (nucleotide accession no.: MN006821, protein ID: QHD44466.1), and the same was confirmed from the Norine database [23,24].
The same sequence was further used for the codon optimization, which was carried out using the Transeq tool. A codon-optimized sequence was further used for the cloning (Figure 1). The protein product translated from the nucleotide sequence was annotated as iturin A, which was considered to be a reference for the further process of cloning and expression through the vector.

3.2. Cloning into the pET-32A Vector

The coding region of the functional gene, iturin A, was amplified from the custom synthesized template DNA. The PCR product revealed a single band of 687 bp, as shown in Figure 2. PCR-amplified products of the gene were purified and cloned into the pET-32A vector and transformed. To confirm the presence of an insert, restriction digestion was performed with the help of BamHI and EcoRI. The transformed colonies were selected based on blue-white screening with colonies grown on ampicillin containing an agar plate. The coding sequences obtained from sequencing were then compared with the reference sequence in the GenBank with its orthologues for confirmation. Several factors, such as high production cost, downstream processing, and lower recovery rate of the biosurfactant, have triggered research on cloning biosurfactant genes to overcome these limitations. Cloning approaches can improvise the biosurfactant secretion capability of strains that can be efficiently utilized for oil recovery from reservoirs. Cloning aids to overcome the time-consuming biosurfactant screening procedures. Researchers have suggested the development and utilization of engineered strains for biosurfactant production with specific applications [25].

3.3. Expression and Purification of His–pET-32A-Iturin A Fusion Protein

The evolution of novel strains and the improvement of metabolites are both aided by genome shuffling. This method was utilized in an attempt to boost iturin A production by optimizing the codon. This approach aided in the high expression of recombinant iturin A, which possesses similar activity as a lipopeptide and surfactin. Hence, a highly expressed iturin A from Bacillus aryabhattai species was developed. The iturin A gene was composed of 687 bp with a conserved acyl transferase domain. The previous development of iturin A required cloning of a ~37 kbp gene cluster responsible for the biosynthesis of iturin A-like lipopeptides. Our attempt was to clone a functional fragment of partial iturin A, resulting in the increased production of the lipopeptide. This technique could be useful for creating microorganisms that operate as biological control agents. The expressed recombinant iturin A was purified on a nickel-NTA agarose column. Purified iturin A was further concentrated to 50 folds with the help of buffer exchange. SDS-PAGE (Figure S1) revealed the purity and molecular mass of iturin A with a fusion protein, which was observed to be 41 kDa (Figure 3). The yield of recombinant iturin A was found to be 60 g/L.
There was a 6.7-fold increase in production in comparison with wild iturin A, which yielded 8.86 g/L under optimized conditions (crude oil, 4.0%; yeast extract, 0.7%; NaNO3, 3.0%), as reported in our previous studies [23]. In a recent study [26], the recombinant strain HZ-PbacA showed high iturin A production owing to the replacement of the promoter PbacA in place of the promoter of the iturin A synthetase cluster. Additionally, an attempt was made to delete the regulator gene abrB to enhance the expression of iturin A synthetase by relieving the repression caused by AbrB on PbacA. As per previous studies, iturin A production from B. subtilis ZK0 and wild-type B. subtilis ZEB01 was found to be 0.005 g/L, which increased to 0.215 g/L by the simultaneous overexpression of the comA and sigA regulatory genes in B. subtilis ZK0 [27]. Dang et al. (2019) [28] reported 0.037 g/L of iturin A when a strong promoter was inserted into the upstream of itu operon for the transcription enhancement of iturin A genes. A further increase in yield (0.113 mg/L) was observed by the overexpression of the DegQ regulator. Several strategies for the overexpression of biosurfactant genes have been reviewed by Jimoh et al. (2021) [29] where the authors have reported the overexpression of signaling molecules that are involved in the regulation of biosynthetic clusters that trigger the production of a biosurfactant. For example, the overexpression of ComX and PhrC in B. subtilis revealed a 6.4-fold improvement in biosurfactant production. A high-level expression of molecules, such as biosurfactants and enzymes, is receiving importance as the functionality attributions of the molecules could be improved through this workflow [30].

3.4. Evaluation of Biosurfactant Activity

Recently, there has been great research interest in biosurfactants due to their exceptional properties, such as specificity, low toxicity, and eco-friendly characteristics. Biosurfactants have a wide range of applications in industries, such as food, medical, pharmaceutical, petrochemical, and agriculture. They have been demonstrated to have the ability to significantly decrease the surface tension and interfacial tension and form microemulsions [10].
The recombinant protein iturin A revealed biosurfactant activities based on a screening assay. The recombinant protein was evaluated for its emulsification index, which was found to be 74 ± 1.9% (Figure 4A). Additionally, the purified recombinant protein showed the dislocation of oil with a 2.2 cm diameter and an oil dispersion area of 29.88 ± 1.47 cm2 during the oil spread assay (Figure 4B). The oil dispersion and significant results of E24 prove the potential biosurfactant nature of recombinant iturin A. An E 24 higher than 30% is usually considered a significant emulsification activity.
In previous studies, a lipopeptide showed 68 ± 0.5% emulsification activity (E24) [7]. A cell-free broth of B. velezensis showed 56.2% E24 with engine oil [31]. A biosurfactant from mutant A. niger M2 showed 62.30% of E24 with 59.81 cm2 oil displacement area [32], and a 66.4 cm2 oil displacement area was shown by a biosurfactant from R. arrhizus [33]. In the present study, recombinant iturin A was revealed to possess potential biosurfactant activity and can be widely employed in oil recovery processes.

3.5. Analytical Characterization

3.5.1. Mass Spectrometry Analysis

The digestion of protein was performed using the trypsin enzyme that cuts at the C-terminal of arginine (R) and lysine (K) [34]. The digestion was performed under two different conditions (protein: trypsin) of 50:1 and 25:1. The fragments were analyzed by MALDI-TOF MS. The MALDI-TOF spectra for both reactions showed the desired digested mass peaks for the protein. The predicted 16 mass fragments were computed with the iturin A primary protein sequence using the Expasy PeptideMass tool (https://web.expasy.org/peptide_mass, accessed on 15 October 2021). The mass spectra (Figure S2) showed 13 fragments for 50:1 reaction (Table 1), and [M + H]+ of iturin A at m/z 213.9814, m/z 627.2845, and m/z 594.3034 was not observed when compared with the predicted fragments, thus indicating 81.25% of protein sequence coverage. On the other hand, 14 fragments for 25:1 reaction were shown (Table 1), indicating 87.5% of the protein sequence coverage (a few lower molecular weight fragments were not observed). Following lower molecular weight fragments, [M + H]+ of iturin A at m/z 627.2845 and m/z 594.3034 was not observed; thus, the MALDI-TOF MS information for iturin A trypsin digestion confirms the identity of the iturin A protein. Previously, several workers reported the molecular masses and varied peaks for different class of peptides based on MALDI-TOF MS spectra. The potential iturin A class comprises different isomers ranging from iturin A2 to A8.
Ye et al. (2012) [35] reported the mass of purified iturin A2 to be 1042. The MALDI-TOF MS analysis of lipopeptides by Labiadh et al. (2021) [36] revealed a varied diversity of molecules, such as bacillomycin D, iturin C, fengycins, and kurstakin from B. subtilis strains. Secretion of lipopeptide isomers, such as iturin (majority), bacillomycin L, and fengycin, was confirmed by Stincone et al. (2020) [37] based on representative peaks (m/z) by mass spectrometry analyses. MALDI-TOF revealed the presence of majority of iturin molecules with fatty acid chain length of C14–C17 and a minor portion of fengycin A and B with a varying fatty acid chain length in the study conducted by Rodríguez-Chávez et al. (2019) [38].

3.5.2. Raman Spectroscopy Analysis

Raman spectroscopic approaches have emerged as useful tools to analyze the composition and secondary structure of pure proteins. Spectral bands in a Raman spectrum are indicative of the presence of specific amino acids, protein structure, side-chain orientations, and local environments. As this method is not sensitive to water, it can be used to characterize materials in solution, preserving their biological activity. The average Raman spectrum shown in Figure 5 has discrete bands representing vibrational modes of the peptide backbone and its side chains. In addition to amide I (1668 cm−1) and III (1266 and 1324 cm−1) bands, amino-acid-specific Raman bands, such as 823 cm−1, 1613 cm−1 (ring breathing modes of tyrosine), 852 cm−1 (proline, hydroxyproline), 935 cm−1 (C–C stretching mode of proline and valine), 1005 cm−1 (phenylalanine), and 1051 cm−1 (C–N stretching), can be clearly seen. Spectral features in the 1050–1060 cm−1 area are suggestive of lipid presence [39] as these bands have been assigned to C–C stretching vibrations. This is further supported by the presence of a strong spectral band at 1466 cm−1 assigned to alkyl CH2 bends present in lipids. Raman spectroscopy has been employed by other researchers for several studies. Shi et al. (2021) [40] studied that a biosurfactant improved the CH4 hydrate structural stability by employing Raman spectra.
From the MALDI MS, the exact masses of the peptide fragments were obtained after trypsin digestion of iturin A. These masses match the predicted tryptic fragments obtained from the Expasy peptide mass bioinformatics tool using the primary sequence of iturin A (https://web.expasy.org/peptide_mass/, accessed on 15 October 2021). The observed masses (from two different tryptic digestion conditions) are within the accepted error range, which confirms the sequence/structure of iturin. On the other hand, from the Raman spectroscopic analysis, characteristic vibration/stretching bands were obtained, which are indicative of specific functional groups, such as amide bonds, certain amino acids (such as Tyr, Pro, and Val), and alkyl/lipid regions that provide preliminary information about the presence of a lipopeptide structure. Although, from the Raman analysis, the exact sequence of the protein cannot be obtained, the functional group identification from Raman data supports the MALDI MS analysis for deducing the structure of the protein based on its primary sequence (obtained from NCBI; https://0-www-ncbi-nlm-nih-gov.brum.beds.ac.uk/, accessed on 15 October 2021). Overall, the techniques are complementary and informative for understanding the structure of a protein [41,42].
Iturin A is a lipopeptide (biosurfactant) that is a cyclic heptapeptide in nature [43] (Figure 6). The amino acid configuration is mainly Asx-Tyr-Asx-(Ser, Glx or Pro)-(Pro or Glu)-(Ser or Asn)-(Thr or Asn) in an LDDLLDL configuration sequence, which is closed by a beta-amino acid linking [44].

3.6. Application of a Biosurfactant for Enhanced Oil Recovery

The application of recombinant iturin A for oil recovery was evaluated a through-sand packed column (Figure S3). The column differs in structure and permeability due to its variation in PV and other parameters. The values of PV, OOIP, and Sorwf ranged from 9.88 to 10.08 mL, 6.92 to 7.18 mL, and 4.49 to 4.59 mL respectively. The residual oil was released by flooding recombinant iturin A into the sand-packed column. The quantity of oil recovered using recombinant iturin A after flooding with brine was 1.55 mL and was used to calculate AOR. Flooding sand-packed columns with recombinant iturin A yielded 61.18 ± 0.85% additional oil, as shown in Table 2. Compared with the negative control, it yielded 29% more oil.
The oil that is trapped in the pores of the sand-packed column imitates the natural oil trapped in the rock pores. Highly viscous residual oil and interfacial tension interferes with the flow properties of oils in wells [45]. The recovery of oil encompasses primary, secondary, and tertiary phases. The natural pressure drive method is used during the primary phase. In the secondary phase, the water flooding process is employed, which leads to the swelling of clay particles and causes subsequent damage to reservoirs. To combat this, brine solution is used during flooding. The application of surfactants is commonly evidenced in the tertiary phase of oil recovery [22]. The addition of a surfactant or biosurfactants exhibits the reduction of interfacial tension among the water and oil layer forming emulsion. It also varies the wettability attributes of the pore surfaces of the reservoir rocks and enhances the mobility of oil [46].
In a similar study conducted by Sharma et al. (2022) [47], 63% recovery of residual oil was observed using a biosurfactant from Franconibacter sp. IITDAS19. The brine solution is employed for enhancing the interactions between the oil and the biosurfactant following the theory of Derjaguin–Landau–Verwey–Overbeek (DLVO). Previously reported studies show that a biosurfactant produced from Bacillus sp. and Fusarium sp. BS-8, when used for oil recovery in a sand-packed column, resulted in 24.5% additional oil recovery [48]. A biosurfactant from Pseudoxanthomonas sp. G3 yielded 20% more oil recovery in comparison with the control in a sand-packed column [21]. Previous studies on oil recovery under simulated conditions reveal the mobilization of 17.15% residual oil by a glycoprotein N-4 biosurfactant [49]. A biosurfactant from Candida tropicalis MTCC 230 has also been reported for enhanced oil recovery [20]. A lipopeptide XT-2 biosurfactant was reported to recover 11–13% of residual oil [50]. The variability in the oil recovery could be due to the different types and concentration of biosurfactants. Recombinant iturin A demonstrated the recovery of a significant portion of residual oil, indicating its potential application in the petroleum industry in the long term.

4. Conclusions

The yield of iturin A was improvised (60 g/L) through the development of a recombinant lipopeptide by cloning, expression, and purification. The molecular mass of purified iturin A was 41 KDa. The recombinant protein was found to be a promising biosurfactant based on the oil spread assay and emulsification index (E 24). The identity of iturin A was confirmed with MALDI-TOF MS, and the structural elucidation of iturin A was carried out by Raman spectroscopy. Recombinant iturin A demonstrated enhanced oil recovery with a sand-packed column. This present study revealed a 6.7-fold increase in recombinant iturin A compared with that of our previous study, and it suggests real-time applications of a recombinant biosurfactant in oil reservoirs and the petroleum industry.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/gels8070403/s1, Figure S1: SDS-PAGE of recombinant iturin A; Figure S2: Maldi MS Spectra for (A) 50:1 trypsin digestion of protein and (B) 25:1 trypsin digestion of recombinant iturin A; Figure S3: Setup for enhanced oil recovery using recombinant iturin A form the sand packed column.

Author Contributions

Z.K.B. and D.A.Y. were involved in the design and conceptualization of the study. D.A.Y. and S.H.D. carried out codon optimization and gene synthesis. D.A.Y., Z.K.B., N.M., C.S., S.H.D., S.D. and D.S. conducted the molecular cloning, expression, and purification part of the experimental work and collected the spectroscopic data. D.A.Y., Z.K.B., N.M., S.P.S. and T.M.Y.K. performed analytical characterization, data analysis, and interpretation. Z.K.B. and D.A.Y. wrote the manuscript. Z.K.B. supervised the overall research work and is involved in resource management. All authors have read and agreed to the published version of the manuscript.

Funding

This work was funded by King Khalid University under grant number R. G. P 2/105/43.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through a research group program under grant number R. G. P 2/105/43. The main corresponding author thanks KLE Technological University, Hubballi, for providing funding to carry out the present research work through research group (RG) projects. The authors would like to acknowledge C. Murali Krishna from ACTREC, Tata Memorial Centre, Mumbai, for facilitating the Raman spectroscopy measurements. We also thank the Mass Spectrometry Facility at the Department of Biological Sciences, Indian Institute of Science, Bangalore, for the MALDI-TOF MS data acquisition.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sekhon, K.; Khanna, S.; Cameotra, S. Enhanced biosurfactant production through cloning of three genes and role of esterase in biosurfactant release. Microb. Cell Fact. 2011, 10, 49. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Kapellos, G.E. Microbial Strategies for Oil Biodegradation. In Modeling of Microscale Transport in Biological Processes; Elsevier: Amsterdam, The Netherlands, 2017; pp. 19–39. [Google Scholar]
  3. Costa, J.A.V.; Treichel, H.; Santos, L.O.; Martins, V.G. Solid-State Fermentation for the Production of Biosurfactants and Their Applications. In Current Developments in Biotechnology and Bioengineering; Elsevier: Amsterdam, The Netherlands, 2018; pp. 357–372. [Google Scholar]
  4. Meena, K.R.; Kanwar, S.S. Lipopeptides as the Antifungal and Antibacterial Agents: Applications in Food Safety and Therapeutics. Biomed Res. Int. 2015, 2015, 473050. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Zohora, U.S.; Rahman, M.S.; Ano, T. Biofilm formation and lipopeptide antibiotic iturin A production in different peptone media. J. Environ. Sci. 2009, 21, S24–S27. [Google Scholar] [CrossRef]
  6. Tsuge, K.; Akiyama, T.; Shoda, M. Cloning, Sequencing, and Characterization of the Iturin A Operon. J. Bacteriol. 2001, 183, 6265–6273. [Google Scholar] [CrossRef] [Green Version]
  7. Yaraguppi, D.A.; Bagewadi, Z.K.; Muddapur, U.M.; Mulla, S.I. Response surface methodology-based optimization of biosurfactant production from isolated Bacillus aryabhattai strain ZDY2. J. Pet. Explor. Prod. Technol. 2020, 10, 2483–2498. [Google Scholar] [CrossRef] [Green Version]
  8. Jimoh, A.A.; Lin, J. Biosurfactant: A new frontier for greener technology and environmental sustainability. Ecotoxicol. Environ. Saf. 2019, 184, 109607. [Google Scholar] [CrossRef]
  9. Bachmann, R.T.; Johnson, A.C.; Edyvean, R.G.J. Biotechnology in the petroleum industry: An overview. Int. Biodeterior. Biodegrad. 2014, 86, 225–237. [Google Scholar] [CrossRef]
  10. Sarubbo, L.A.; Silva, M.d.G.C.; Durval, I.J.B.; Bezerra, K.G.O.; Ribeiro, B.G.; Silva, I.A.; Twigg, M.S.; Banat, I.M. Biosurfactants: Production, properties, applications, trends, and general perspectives. Biochem. Eng. J. 2022, 181, 108377. [Google Scholar] [CrossRef]
  11. Roongsawang, N.; Thaniyavarn, J.; Thaniyavarn, S.; Kameyama, T.; Haruki, M.; Imanaka, T.; Morikawa, M.; Kanaya, S. Isolation and characterization of a halotolerant Bacillus subtilis BBK-1 which produces three kinds of lipopeptides: Bacillomycin L, plipastatin, and surfactin. Extremophiles 2002, 6, 499–506. [Google Scholar] [CrossRef]
  12. Luo, C.; Chen, Y.; Liu, X.; Wang, X.; Wang, X.; Li, X.; Zhao, Y.; Wei, L. Engineered biosynthesis of cyclic lipopeptide locillomycins in surrogate host Bacillus velezensis FZB42 and derivative strains enhance antibacterial activity. Appl. Microbiol. Biotechnol. 2019, 103, 4467–4481. [Google Scholar] [CrossRef]
  13. Ebadi, A.; Olamaee, M.; Khoshkholgh Sima, N.A.; Ghorbani Nasrabadi, R.; Hashemi, M. Isolation and Characterization of Biosurfactant Producing and Crude Oil Degrading Bacteria from Oil Contaminated Soils. Iran. J. Sci. Technol. Trans. A Sci. 2018, 42, 1149–1156. [Google Scholar] [CrossRef]
  14. Yaraguppi, D.A.; Bagewadi, Z.K.; Deshpande, S.H.; Chandramohan, V. In Silico Study on the Inhibition of UDP-N-Acetylglucosamine 1-Carboxy Vinyl Transferase from Salmonella typhimurium by the Lipopeptide Produced from Bacillus aryabhattai. Int. J. Pept. Res. Ther. 2022, 28, 80. [Google Scholar] [CrossRef]
  15. Duarte, C.; Gudiña, E.J.; Lima, C.F.; Rodrigues, L.R. Effects of biosurfactants on the viability and proliferation of human breast cancer cells. AMB Express 2014, 4, 40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Chen, M.; Xu, P.; Zeng, G.; Yang, C.; Huang, D.; Zhang, J. Bioremediation of soils contaminated with polycyclic aromatic hydrocarbons, petroleum, pesticides, chlorophenols and heavy metals by composting: Applications, microbes and future research needs. Biotechnol. Adv. 2015, 33, 745–755. [Google Scholar] [CrossRef] [PubMed]
  17. Shi, J.; Zhu, X.; Lu, Y.; Zhao, H.; Lu, F.; Lu, Z. Improving Iturin a Production of Bacillus amyloliquefaciens by Genome Shuffling and Its Inhibition against Saccharomyces cerevisiae in Orange Juice. Front. Microbiol. 2018, 9, 2683. [Google Scholar] [CrossRef]
  18. Yaish, M.W. Draft Genome Sequence of the Endophytic Bacillus aryabhattai Strain SQU-R12, Identified from Phoenix dactylifera L. Roots. Genome Announc. 2017, 5, e00718-17. [Google Scholar] [CrossRef] [Green Version]
  19. Madeira, F.; Park, Y.m.; Lee, J.; Buso, N.; Gur, T.; Madhusoodanan, N.; Basutkar, P.; Tivey, A.R.N.; Potter, S.C.; Finn, R.D.; et al. The EMBL-EBI search and sequence analysis tools APIs in 2019. Nucleic Acids Res. 2019, 47, W636–W641. [Google Scholar] [CrossRef] [Green Version]
  20. Ashish; Debnath, M. Application of biosurfactant produced by an adaptive strain of C.tropicalis MTCC230 in microbial enhanced oil recovery (MEOR) and removal of motor oil from contaminated sand and water. J. Pet. Sci. Eng. 2018, 170, 40–48. [Google Scholar] [CrossRef]
  21. Astuti, D.I.; Purwasena, I.A.; Putri, R.E.; Amaniyah, M.; Sugai, Y. Screening and characterization of biosurfactant produced by Pseudoxanthomonas sp. G3 and its applicability for enhanced oil recovery. J. Pet. Explor. Prod. Technol. 2019, 9, 2279–2289. [Google Scholar] [CrossRef] [Green Version]
  22. Das, A.J.; Kumar, R. Utilization of agro-industrial waste for biosurfactant production under submerged fermentation and its application in oil recovery from sand matrix. Bioresour. Technol. 2018, 260, 233–240. [Google Scholar] [CrossRef]
  23. Yaraguppi, D.A.; Deshpande, S.H.; Bagewadi, Z.K.; Kumar, S.; Muddapur, U.M. Genome Analysis of Bacillus aryabhattai to Identify Biosynthetic Gene Clusters and In Silico Methods to Elucidate its Antimicrobial Nature. Int. J. Pept. Res. Ther. 2021, 27, 1331–1342. [Google Scholar] [CrossRef]
  24. Flissi, A.; Ricart, E.; Campart, C.; Chevalier, M.; Dufresne, Y.; Michalik, J.; Jacques, P.; Flahaut, C.; Lisacek, F.; Leclère, V.; et al. Norine: Update of the nonribosomal peptide resource. Nucleic Acids Res. 2020, 48, D465–D469. [Google Scholar] [CrossRef]
  25. Gaur, V.K.; Sharma, P.; Gupta, S.; Varjani, S.; Srivastava, J.K.; Wong, J.W.C.; Ngo, H.H. Opportunities and challenges in omics approaches for biosurfactant production and feasibility of site remediation: Strategies and advancements. Environ. Technol. Innov. 2022, 25, 102132. [Google Scholar] [CrossRef]
  26. Xu, Y.; Cai, D.; Zhang, H.; Gao, L.; Yang, Y.; Gao, J.; Li, Y.; Yang, C.; Ji, Z.; Yu, J.; et al. Enhanced production of iturin A in Bacillus amyloliquefaciens by genetic engineering and medium optimization. Process Biochem. 2020, 90, 50–57. [Google Scholar] [CrossRef]
  27. Zhang, Z.; Ding, Z.T.; Zhong, J.; Zhou, J.Y.; Shu, D.; Luo, D.; Yang, J.; Tan, H. Improvement of iturin A production in Bacillus subtilis ZK0 by overexpression of the comA and sigA genes. Lett. Appl. Microbiol. 2017, 64, 452–458. [Google Scholar] [CrossRef] [PubMed]
  28. Dang, Y.; Zhao, F.; Liu, X.; Fan, X.; Huang, R.; Gao, W.; Wang, S.; Yang, C. Enhanced production of antifungal lipopeptide iturin A by Bacillus amyloliquefaciens LL3 through metabolic engineering and culture conditions optimization. Microb. Cell Fact. 2019, 18, 68. [Google Scholar] [CrossRef] [PubMed]
  29. Jimoh, A.A.; Senbadejo, T.Y.; Adeleke, R.; Lin, J. Development and Genetic Engineering of Hyper-Producing Microbial Strains for Improved Synthesis of Biosurfactants. Mol. Biotechnol. 2021, 63, 267–288. [Google Scholar] [CrossRef]
  30. Bagewadi, Z.K.; Mulla, S.I.; Ninnekar, H.Z. Purification, characterization, gene cloning and expression of GH-10 xylanase (Penicillium citrinum isolate HZN13). 3 Biotech 2016, 6, 169. [Google Scholar] [CrossRef] [Green Version]
  31. Meena, K.R.; Dhiman, R.; Singh, K.; Kumar, S.; Sharma, A.; Kanwar, S.S.; Mondal, R.; Das, S.; Franco, O.L.; Mandal, A.K. Purification and identification of a surfactin biosurfactant and engine oil degradation by Bacillus velezensis KLP2016. Microb. Cell Fact. 2021, 20, 26. [Google Scholar] [CrossRef]
  32. Asgher, M.; Arshad, S.; Qamar, S.A.; Khalid, N. Improved biosurfactant production from Aspergillus niger through chemical mutagenesis: Characterization and RSM optimization. SN Appl. Sci. 2020, 2, 966. [Google Scholar] [CrossRef] [Green Version]
  33. Pele, M.A.; Ribeaux, D.R.; Vieira, E.R.; Souza, A.F.; Luna, M.A.C.; Rodríguez, D.M.; Andrade, R.F.S.; Alviano, D.S.; Alviano, C.S.; Barreto-Bergter, E.; et al. Conversion of renewable substrates for biosurfactant production by Rhizopus arrhizus UCP 1607 and enhancing the removal of diesel oil from marine soil. Electron. J. Biotechnol. 2019, 38, 40–48. [Google Scholar] [CrossRef]
  34. Simpson, R.J. Fragmentation of Protein Using Trypsin. Cold Spring Harb. Protoc. 2006, 2006, pdb.prot4550. [Google Scholar] [CrossRef] [PubMed]
  35. Ye, Y.; Li, Q.; Fu, G.; Yuan, G.; Miao, J.; Lin, W. Identification of Antifungal Substance (Iturin A2) Produced by Bacillus subtilis B47 and its Effect on Southern Corn Leaf Blight. J. Integr. Agric. 2012, 11, 90–99. [Google Scholar] [CrossRef]
  36. Labiadh, M.; Dhaouadi, S.; Chollet, M.; Chataigne, G.; Tricot, C.; Jacques, P.; Flahaut, S.; Kallel, S. Antifungal lipopeptides from Bacillus strains isolated from rhizosphere of Citrus trees. Rhizosphere 2021, 19, 100399. [Google Scholar] [CrossRef]
  37. Stincone, P.; Veras, F.F.; Pereira, J.Q.; Mayer, F.Q.; Varela, A.P.M.; Brandelli, A. Diversity of cyclic antimicrobial lipopeptides from Bacillus P34 revealed by functional annotation and comparative genome analysis. Microbiol. Res. 2020, 238, 126515. [Google Scholar] [CrossRef] [PubMed]
  38. Rodríguez-Chávez, J.L.; Juárez-Campusano, Y.S.; Delgado, G.; Pacheco Aguilar, J.R. Identification of lipopeptides from Bacillus strain Q11 with ability to inhibit the germination of Penicillium expansum, the etiological agent of postharvest blue mold disease. Postharvest Biol. Technol. 2019, 155, 72–79. [Google Scholar] [CrossRef]
  39. San Keskin, N.O.; Han, D.; Devrim Ozkan, A.; Angun, P.; Onarman Umu, O.C.; Tekinay, T. Production and structural characterization of biosurfactant produced by newly isolated staphylococcus xylosus STF1 from petroleum contaminated soil. J. Pet. Sci. Eng. 2015, 133, 689–694. [Google Scholar] [CrossRef]
  40. Shi, L.; He, Y.; Lu, J.; Hou, G.; Liang, D. Anti-agglomeration evaluation and Raman spectroscopic analysis on mixed biosurfactants for preventing CH4 hydrate blockage in n-octane + water systems. Energy 2021, 229, 120755. [Google Scholar] [CrossRef]
  41. Zhu, F.; Isaacs, N.W.; Hecht, L.; Barron, L.D. Raman Optical Activity: A Tool for Protein Structure Analysis. Structure 2005, 13, 1409–1419. [Google Scholar] [CrossRef]
  42. Maiti, N.C.; Apetri, M.M.; Zagorski, M.G.; Carey, P.R.; Anderson, V.E. Raman Spectroscopic Characterization of Secondary Structure in Natively Unfolded Proteins: α-Synuclein. J. Am. Chem. Soc. 2004, 126, 2399–2408. [Google Scholar] [CrossRef]
  43. Elsevier. Biobased Surfactants; Elsevier: Amsterdam, The Netherlands, 2019; ISBN 9780128127056. [Google Scholar]
  44. Hashizume, H.; Nishimura, Y. Cyclic Lipopeptide Antibiotics. Stud. Nat. Prod. Chem. 2008, 35, 693–751. [Google Scholar]
  45. Pandey, R.; Krishnamurthy, B.; Singh, H.P.; Batish, D.R. Evaluation of a glycolipopepetide biosurfactant from Aeromonas hydrophila RP1 for bioremediation and enhanced oil recovery. J. Clean. Prod. 2022, 345, 131098. [Google Scholar] [CrossRef]
  46. Nasiri, M.A.; Biria, D. Extraction of the indigenous crude oil dissolved biosurfactants and their potential in enhanced oil recovery. Colloids Surf. A Physicochem. Eng. Asp. 2020, 603, 125216. [Google Scholar] [CrossRef]
  47. Sharma, J.; Kapley, A.; Sundar, D.; Srivastava, P. Characterization of a potent biosurfactant produced from Franconibacter sp. IITDAS19 and its application in enhanced oil recovery. Colloids Surf. B Biointerfaces 2022, 214, 112453. [Google Scholar] [CrossRef] [PubMed]
  48. Suthar, H.; Hingurao, K.; Desai, A.; Nerurkar, A. Evaluation of bioemulsifier mediated Microbial Enhanced Oil Recovery using sand pack column. J. Microbiol. Methods 2008, 75, 225–230. [Google Scholar] [CrossRef]
  49. Arora, P.; Kshirsagar, P.R.; Rana, D.P.; Dhakephalkar, P.K. Hyperthermophilic Clostridium sp. N-4 produced a glycoprotein biosurfactant that enhanced recovery of residual oil at 96 °C in lab studies. Colloids Surf. B Biointerfaces 2019, 182, 110372. [Google Scholar] [CrossRef]
  50. Wang, X.-T.; Liu, B.; Li, X.-Z.; Lin, W.; Li, D.-A.; Dong, H.; Wang, L. Biosurfactants produced by novel facultative-halophilic Bacillus sp. XT-2 with biodegradation of long chain n-alkane and the application for enhancing waxy oil recovery. Energy 2022, 240, 122802. [Google Scholar] [CrossRef]
Figure 1. Graphical representation of the lpa-14 (iturin synthesis gene) track with protein features and region features. Sfp-4′-phosphopantetheinyl transferase; QHD44466.1—protein ID (antimicrobial lipopeptide (iturin).
Figure 1. Graphical representation of the lpa-14 (iturin synthesis gene) track with protein features and region features. Sfp-4′-phosphopantetheinyl transferase; QHD44466.1—protein ID (antimicrobial lipopeptide (iturin).
Gels 08 00403 g001
Figure 2. Cloning of the lpa-14 gene from Bacillus aryabhattai: (A) amplification of the iturin A gene by PCR and (B) recombinant plasmid showing a band size of 687 bp (iturin A gene) and a band size of 6.5 Kb (pET-32A vector).
Figure 2. Cloning of the lpa-14 gene from Bacillus aryabhattai: (A) amplification of the iturin A gene by PCR and (B) recombinant plasmid showing a band size of 687 bp (iturin A gene) and a band size of 6.5 Kb (pET-32A vector).
Gels 08 00403 g002
Figure 3. SDS-PAGE of recombinant iturin A. (Lane 1) Protein sample, (Lane 2) protein mol. wt marker (25 to 250 kDa), (Lane 3): protein sample.
Figure 3. SDS-PAGE of recombinant iturin A. (Lane 1) Protein sample, (Lane 2) protein mol. wt marker (25 to 250 kDa), (Lane 3): protein sample.
Gels 08 00403 g003
Figure 4. (A) Emulsification activity of recombinant iturin A: (i) test sample, (ii) negative control. (B) Oil spread assay: (i) motor oil before adding recombinant lipopeptide, (ii) recombinant lipopeptide added on oil surface, (iii) oil dispersion by recombinant lipopeptide.
Figure 4. (A) Emulsification activity of recombinant iturin A: (i) test sample, (ii) negative control. (B) Oil spread assay: (i) motor oil before adding recombinant lipopeptide, (ii) recombinant lipopeptide added on oil surface, (iii) oil dispersion by recombinant lipopeptide.
Gels 08 00403 g004
Figure 5. Mean Raman spectrum of purified recombinant iturin A.
Figure 5. Mean Raman spectrum of purified recombinant iturin A.
Gels 08 00403 g005
Figure 6. Two-dimensional structural representation of a biosurfactant (iturin A).
Figure 6. Two-dimensional structural representation of a biosurfactant (iturin A).
Gels 08 00403 g006
Table 1. The predicted mass fragments and the observed mass fragments for 50:1 and 25:1 recombinant iturin A trypsin digestion reaction.
Table 1. The predicted mass fragments and the observed mass fragments for 50:1 and 25:1 recombinant iturin A trypsin digestion reaction.
Predicted Masses, [M + H]+Observed Masses in 50:1 Trypsin Digestion, [M + H]+Observed Masses in 25:1 Trypsin Digestion, [M + H]+
3904.93904.23904.9
2135.9-2135.9
2095.92096.02095.9
1970.91970.91970.9
1886.01886.01886.0
1819.81819.81819.8
1532.71532.71532.7
1290.71290.81290.7
1133.41133.61133.5
1073.61073.61073.5
1031.61031.61031.6
937.4937.4937.4
767.3767.4767.4
627.3--
623.3623.3623.3
594.3--
Table 2. Enhanced oil recovery from the sand-packed column using recombinant iturin A.
Table 2. Enhanced oil recovery from the sand-packed column using recombinant iturin A.
ParametersSamplePositive ControlNegative Control
Pore volume (mL)10 ± 0.89.97 ± 0.0910 ± 0.08
Porosity (%)20 ± 0.1619.93 ± 0.1820 ± 00.16
OOIP (mL)7.1± 0.087 ± 0.087.03 ± 0.08
Soi (%)71 ± 0.7370.23 ± 0.4170.34 ± 1.45
Swi (%)29 ± 0.7329.76 ± 0.4129.65 ± 0.46
Sorwf (mL)4.53 ± 0.044.55 ± 0.044.52 ± 0.02
Sor (%)35.85 ± 0.5835 ± 0.6935.77 ± 0.75
Sorif (mL)1.51 ± 0.041.05 ± 0.050.80 ± 0.03
AOR (%)61.18 ± 0.8542.85 ± 0.8432.06 ± 1.04
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Yaraguppi, D.A.; Bagewadi, Z.K.; Mahanta, N.; Singh, S.P.; Khan, T.M.Y.; Deshpande, S.H.; Soratur, C.; Das, S.; Saikia, D. Gene Expression and Characterization of Iturin A Lipopeptide Biosurfactant from Bacillus aryabhattai for Enhanced Oil Recovery. Gels 2022, 8, 403. https://0-doi-org.brum.beds.ac.uk/10.3390/gels8070403

AMA Style

Yaraguppi DA, Bagewadi ZK, Mahanta N, Singh SP, Khan TMY, Deshpande SH, Soratur C, Das S, Saikia D. Gene Expression and Characterization of Iturin A Lipopeptide Biosurfactant from Bacillus aryabhattai for Enhanced Oil Recovery. Gels. 2022; 8(7):403. https://0-doi-org.brum.beds.ac.uk/10.3390/gels8070403

Chicago/Turabian Style

Yaraguppi, Deepak A., Zabin K. Bagewadi, Nilkamal Mahanta, Surya P. Singh, T. M. Yunus Khan, Sanjay H. Deshpande, Chaitra Soratur, Simita Das, and Dimple Saikia. 2022. "Gene Expression and Characterization of Iturin A Lipopeptide Biosurfactant from Bacillus aryabhattai for Enhanced Oil Recovery" Gels 8, no. 7: 403. https://0-doi-org.brum.beds.ac.uk/10.3390/gels8070403

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop