26
Accepted Manuscript Title: Identification and characterization of biosurfactants produced by the Arctic bacterium Pseudomonas putida BD2 Author: Tomasz Janek Marcin Lukaszewicz Anna Krasowska PII: S0927-7765(13)00318-4 DOI: http://dx.doi.org/doi:10.1016/j.colsurfb.2013.05.008 Reference: COLSUB 5788 To appear in: Colloids and Surfaces B: Biointerfaces Received date: 12-12-2012 Revised date: 23-4-2013 Accepted date: 7-5-2013 Please cite this article as: T. Janek, M. Lukaszewicz, A. Krasowska, Identification and characterization of biosurfactants produced by the Arctic bacterium Pseudomonas putida BD2, Colloids and Surfaces B: Biointerfaces (2013), http://dx.doi.org/10.1016/j.colsurfb.2013.05.008 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Identification and characterization of biosurfactants

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Identification and characterization of biosurfactants

Accepted Manuscript

Title: Identification and characterization of biosurfactantsproduced by the Arctic bacterium Pseudomonas putida BD2

Author: Tomasz Janek Marcin Łukaszewicz Anna Krasowska

PII: S0927-7765(13)00318-4DOI: http://dx.doi.org/doi:10.1016/j.colsurfb.2013.05.008Reference: COLSUB 5788

To appear in: Colloids and Surfaces B: Biointerfaces

Received date: 12-12-2012Revised date: 23-4-2013Accepted date: 7-5-2013

Please cite this article as: T. Janek, M. Łukaszewicz, A. Krasowska, Identificationand characterization of biosurfactants produced by the Arctic bacteriumPseudomonas putida BD2, Colloids and Surfaces B: Biointerfaces (2013),http://dx.doi.org/10.1016/j.colsurfb.2013.05.008

This is a PDF file of an unedited manuscript that has been accepted for publication.As a service to our customers we are providing this early version of the manuscript.The manuscript will undergo copyediting, typesetting, and review of the resulting proofbefore it is published in its final form. Please note that during the production processerrors may be discovered which could affect the content, and all legal disclaimers thatapply to the journal pertain.

Page 2: Identification and characterization of biosurfactants

Page 1 of 25

Accep

ted

Man

uscr

ipt

1

Identification and characterization of biosurfactants produced by the Arctic bacterium

Pseudomonas putida BD2

Tomasz Janeka,*, Marcin Łukaszewicza,b, Anna Krasowskaa

a) Department of Biotransformation, Faculty of Biotechnology, University of Wroclaw,

Przybyszewskiego 63/77, 51-148 Wrocław, Poland

b) Faculty of Chemistry, Wroclaw University of Technology, Gdańska 7/9, 50-344 Wrocław,

Poland

Abstract

One hundred and thirty bacterial strains, isolated from Arctic soil on the Svalbard

Archipelago, were screened for biosurfactant production. Among them, an isolate identified

as Pseudomonas putida BD2 was selected as a potential biosurfactant-producer based on the

surface/interfacial activity of the culture supernatant. The ability of the strain to produce

simultaneously phosphatidylethanolamines and rhamnolipid, using glucose as a sole carbon

source, was demonstrated. The rhamnolipid Rha-Rha-C10-C10 and two homologs of

phosphatidylethanolamine were extracted from cell-free supernatant of P. putida BD2 culture

with ethyl acetate and identified by UPLC-MS analysis. For Rha-Rha-C10-C10 the surface

tension decreased from 72 to 31 mN/m and the critical micelle concentration was 0.130

mg/mL. The Rha-Rha-C10-C10 was able to form stable aggregates (80-121 nm). Pretreatment

of a polystyrene surface with 0.5 mg/mL rhamnolipid inhibited bacterial adhesion by 43-79%

and that of the pathogenic fungal species C. albicans by 89-90%. The same concentration of

phosphatidylethanolamines inhibited bacterial adhesion by 23-72% and that of C. albicans by

96%- 98%. To our knowledge, this is the first report where one type rhamnolipid and two

homologs of phospholipid biosurfactants were produced by Pseudomonas putida isolated

from Arctic soil.

Keywords: Pseudomonas putida BD2; glycolipid; phosphatidylethanolamine; biosurfactant

* Corresponding author. Tel.: +48 713756210; fax: +48 713756234

E-mail address: [email protected] (Tomasz Janek).

Page 3: Identification and characterization of biosurfactants

Page 2 of 25

Accep

ted

Man

uscr

ipt

2

1. Introduction

Biosurfactants are a structurally diverse group of surface-active compounds

synthesized by microorganisms [1, 2]. The structure of these secondary metabolites is strictly

dependent on the medium composition, e.g. carbon and nitrogen sources. Microbial

surfactants encompass a wide spectrum of molecules, such as lipopeptides, glycolipids,

phospholipids, fatty acids, neutral lipids and polymeric biosurfactants [3]. Microorganisms

such as bacteria, filamentous fungi and yeast produce biosurfactants extracellularly or as

compounds associated with cell membrane [4].

These natural compounds have many advantages over synthetic surfactants, such as

non-toxic character or high biodegradability, and have extensive applications in many

industrial fields such as production of food, cosmetics, pesticides, detergents,

pharmaceuticals, oil recovery and bioremediation [5, 6]. Besides their potential application in

industrial and environmental remediation, these surface-active compounds have been reported

to possess several properties of therapeutic and biomedical importance. The features of

biosurfactants allow for, e.g. antimicrobial and anti-adhesive action against several pathogenic

microorganisms on medical implants [7, 8] restoration and maintenance of bacterial

homeostasis in human body [9] or antibiotic activity against bacteria and fungi [10].

Glycolipids are one of the most promising biosurfactants. They consist of

carbohydrates and long-chain aliphatic or hydroxyaliphatic acids. Rhamnolipids (RLs)

produced by Pseudomonas strains have a glycosyl head group (a rhamnose moiety) and a 3-

(hydroxyalkanoyloxy)alkanoic acid (HAA) [11]. The two most abundant RLs congeners are

mono-rhamnolipids and di-rhamnolipids often referred to as Rha-C10-C10 and Rha-Rha-C10-

C10, respectively [12]. Factors influencing RLs production in the genus Pseudomonas were

reviewed by Muller et al. [13]. Pseudomonas was shown to be capable of using both water

soluble carbon substrates, such as glycerol, mannitol and fructose [14] as well as water

immiscible n-alkanes to produce rhamnolipid-type biosurfactants [15]. Not only the type of

carbon and nitrogen source but also the respective C/N ratio strongly influence total RL

production [16, 17].

Desai and Banat [1] have reviewed a wide range of techniques to determine the

presence of biosurfactants in culture media. The chemical and physico-chemical methods

used for this purpose include determination of surface tension (ST), emulsification index (E24)

determined over a 24-h period [18], drop-collapsing test [19], thin layer chromatography

Page 4: Identification and characterization of biosurfactants

Page 3 of 25

Accep

ted

Man

uscr

ipt

3

(TLC), high performance liquid chromatography (HPLC), Fourier transform infrared

spectroscopy (FTIR) and mass spectrometry (MS).

The aims of this work included identification of a new Arctic strain, purification of

biosurfactants secreted to minimal medium with glucose as a sole carbon source,

characterization of chemical structures of isolated biosurfactants and their physico-chemical

properties (minimum surface tension, critical micelle concentration, emulsification activity)

and the anti-adhesive activities against several microorganisms. Furthermore, the size of

micelles of Rha-Rha-C10-C10 was also studied.

2. Experimental

2.1. Microorganism and growth culture

Pseudomonas putida BD2 was isolated from a soil sample from the Arctic

Archipelago of Svalbard (latitude 77o 04' N, longitude 15o 14' E). After being grown on LB

medium containing 5 g/L yeast extract, 10 g/L bacto-tryptone and 10 g/L NaCl, the bacterial

strain was maintained at −70 °C in 15% glycerol. Experiments were carried out in 1000 mL

baffled Erlenmeyer flasks containing 500 mL of medium with the following composition: 7

g/L K2HPO4, 2 g/L KH2PO4, 1 g/L (NH4)2SO4, 0.5 g/L sodium citrate x 2H2O, and 0.1 g/L

MgSO4 x 7H2O (pH 7.0) [20]. Finally, 20 g/L of glucose was used as the carbon source.

Medium components were sterilized separately at 120 °C, 1 atm for 20 min. The flasks were

inoculated with 5 ml overnight pre-culture of the strain and incubated at 28 °C on a reciprocal

rotary shaker at 170 rpm.

2.2. Identification of bacterial strain

Strain BD2 was identified by the API 20E test for Enterobacteriacae (BioMerieux,

Marcy l’Etoile, France) and genomic DNA was obtained by the method of Janek et al. [20].

Genomic DNA was isolated from a one-day liquid culture using a GeneMATRIX Bacterial

and Yeast Genomic DNA Purification kit 50 (EURx, Gdansk, Poland) following the

manufacturer’s standard protocol, and used as a target for polymerase chain reaction (PCR)

amplification using primers 27F (5’-AGR GTT YGA TYM TGG CTC AG-3’) and 1492R

(5’-GGT TAC CTT GTT ACG ACT T-3’). The PCR products were subjected to agarose gel

electrophoresis, purified using a GeneMATRIX PCR/DNA Clean-up Purification kits (EURx,

Gdansk, Poland), and their nucleotide sequence was determined at the Institute of

Biochemistry and Biophysics, Polish Academy of Sciences (Warsaw, Poland) using the

Page 5: Identification and characterization of biosurfactants

Page 4 of 25

Accep

ted

Man

uscr

ipt

4

standard shotgun sequencing reagents and a 454 GS FLX Titanium Sequencing System

(Roche, Basel, Switzerland), according to the manufacturer's instructions. The rDNA gene

sequence obtained from the BD2 isolate was compared with other bacterial sequences by

using GenBank (http://www.ncbi.nlm.nih.gov/BLAST/).

2.3. Biosurfactant production and purification

Microbial growth was evaluated by measuring the absorbance of the culture at 600

nm. Glucose released was measured enzymatically with the Glucose kit from Biosystems

(Barcelona, Spain). Emulsification assays of the biosurfactants were performed using a

previously described method [18]. The emulsification activity of the supernatant was

measured by adding 6 mL petroleum ether to 4 ml of aqueous sample in a test tube, vortexing

for 2 min, and then leaving it to settle for 24 h. The emulsification index (E24) was estimated

as the height of the emulsion layer, divided by the total height, multiplied by 100. The surface

tension of the free culture supernatant was measured with a tensiometer using the du Nouy

procedure with a platinum ring at 25 °C.

Cell-free supernatant was obtained from the culture by centrifugation (8000g) at 4 °C

for 30 min. It was extracted three times with ethyl acetate and evaporated under vacuum. The

crude biosurfactants were dissolved in 5 mL of methanol, filtered through a 0.2 µm syringe

filter and 40 mg crude biosurfactants were separated by preparative layer chromatography

(PLC) (silica gel 60 F254 20 cm x 20 cm with concentrating zone, 1 mm x 20 cm x 4 cm,

Merck; chloroform/methanol/water 65:15:2). Visualization was carried out by UV

transilluminator. In preparative mode, visualized spots were scraped off and the biosurfactants

were extracted with 15 mL of chloroform/methanol (2:1). The surface-active fractions were

evaporated under vacuum and stored at -20 °C for further studies.

2.4. Thin-layer chromatography

The purified biosurfactant fractions obtained after preparative TLC purification were

separated by thin layer chromatography (Silica gel 60; Merck, Darmstadt, Germany). The

compounds were separated using a mobile phase of chloroform/methanol/water in a 65:15:2

ratio (vol/vol/vol). The resulting spots on the TLC were visualized by spraying with a solution

of 0.25% ninhydrin in acetone [20] for amine groups and orcinol solutions [8] for sugar

detection. For the detection of lipids, the plates were treated with 0.1% bromothymol blue in

10% aqueous ethanol [21] and iodine.

Page 6: Identification and characterization of biosurfactants

Page 5 of 25

Accep

ted

Man

uscr

ipt

5

2.5. Liquid chromatography

Aliquots of the biosurfactant extracts or of the chromatographed samples were

dissolved in methanol to obtain 1 mg/mL solutions.

Chromatographic equipment consisted of an Acquity ultra-performance liquid

chromatography (UPLC) system (Waters, Milford, MA, USA). Separation was performed on

a BEH C18 column (1.7 µm, 2.1 x 100 mm, Waters) held at 40 oC. A multistep linear gradient

composed of eluent A (water + 0.1% trifluoroacetic acid) and eluent B (acetonitrile + 0.1%

trifluoroacetic acid) was applied. The autosampler temperature was maintained at 10 oC and

10 µL of sample solution was injected. From 0.00 min to 13.00 min a linear gradient was

applied from the mixture A:B (70:30, vol/vol) to A:B (0:100, vol/vol). A plateau of 100%

eluent B from 13.00 min to 15.00 min was set before going back to 70% eluent A from 15.00

min to 16.00 min. Flow rate was 0.3 mL/min.

2.5. Mass spectrometry analysis

The LC system was coupled to a Waters Xevo QTOF MS mass spectrometer with an

atmospheric pressure electrospray interface. The ESI source was set in positive and negative

ionization mode. The parameters used for the mass spectrometer under the ESI mode were as

follows: ESI+ (capillary voltage 4.50 kV, source temperature 80 °C, desolvation temperature

250 °C, desolvation gas flow 650 L/h, cone gas flow 18 L/h) and ESI- (capillary voltage 3.50

kV, source temperature 120 °C, desolvation temperature 250 °C, desolvation gas flow 700 L/h,

cone gas flow 20 L/h). Different cone voltages were tested to study the fragmentation of

biosurfactants. LC/MS full scan positive and negative modes were performed from m/z 200 to

800; alternatively, LC/ESI-MS/MS modalities were applied to the selected precursor ions,

following the conditions set during the infusion analysis. Waters MassLynx version NT 4.1

was used for LC/MS system control and data analysis.

2.6. Physico-chemical characterization

The surface tension was measured at 25 °C with a Kruss K100 (Kruss GmbH,

Hamburg, Germany), tensiometer by the du Nouy’s ring method. The instrument was

calibrated against Mili-Q ultrapure distilled water. Aqueous solutions of purified rhamnolipids

in the concentration range of 180–5 mg/L were obtained by successive dilutions of a

concentrated sample prepared by weight in ultrapure water.

Page 7: Identification and characterization of biosurfactants

Page 6 of 25

Accep

ted

Man

uscr

ipt

6

The mean particle size and polydispersity index (PDI) of rhamnolipids diluted in

Milli-Q water were determined using a Zetasizer Nano-ZS (Malvern Instruments Ltd.,

Malvern, UK) and PCS software. The analysis of particle size and PDI, determined by photon

correlation spectroscopy, was done using the volume distribution algorithm. The

polydispersity index qualifies the particle size distribution, which here ranged from 0 for

monodisperse to 1.0 for entirely heterodisperse emulsions. In order to obtain the optimum

light scattering intensity and for the size analysis, approximately 100 μL of the rhamnolipid

was added to 900 µL of Milli-Q water. All the measurements were carried out at 25 ºC.

Emulsification activity of phospholipids and rhamnolipids was tested on different

organic compounds: petroleum ether, benzene, n-hexane, cyclohexane, n-hexadecane, xylene,

toluene, sunflower, olive and rapeseed oil. The purified biosurfactants dissolved in 5.0 mL

distilled water (0.2% w/v) were mixed with 5 mL of each hydrophobic compound, and then

vortexed at high speed for 2 min, after which the mixture was kept at 25 °C for 24 h.

2.7. Biological assays

The anti-adhesive properties of phospholipids and rhamnolipids were tested on several

pathogenic strains that colonize medical devices or human body. A wide range of Gram-

positive and Gram-negative bacteria were tested: Escherichia coli ATCC 25922, Escherichia

coli ATCC 10536, Escherichia coli 17-2 (clinical isolate, Wroclaw Medical University),

Enterococcus faecalis ATCC 29212, Enterococcus faecalis JA/3 (clinical isolate, Wroclaw

Medical University), Eenterococcus hirae ATCC 10541, Staphylococcus epidermidis KCTC

1917 and Proteus mirabilis ATCC 21100. Yeast strains: Candida albicans ATCC 20231,

Candida albicans SC5314.

Inhibition of microbial adhesion by the biosurfactants was tested in 96-well plates

(Sarstedt, Nümbrecht, Germany) by the method of Janek et al. [7]. The wells of a sterile 96-

well flat-bottom plate were filled with biosurfactants dissolved in PBS. Next, the plates were

incubated for 2 h at 37 °C on a rotary shaker (MixMate, Eppendorf, Hamburg, Germany) at

300 rpm and subsequently washed twice with PBS. The overnight cultures of microbial

strains were centrifuged, washed twice with PBS (pH 7.4) and re-suspended in PBS. The

microbial suspension was added to each well of the microtiter plate. After a 2-h incubation at

37 °C in a rotary shaker at 300 rpm, nonadherent cells were removed by three washes with

PBS. Then the plates were stained with 0.1% crystal violet for 5 min and again washed three

times with PBS. The adherent microorganisms were permeabilized and the dye was

resolubilized with 150 μL of isopropanol-0.04 N HCl and 50 μL of 0.25% SDS per well.

Page 8: Identification and characterization of biosurfactants

Page 7 of 25

Accep

ted

Man

uscr

ipt

7

Crystal violet optical density readings of each well were taken at 590 nm on the Asys UVM

340 (Biogenet) microplate. All the experiments were carried out in triplicate with suitable

controls (uninoculated microtiter plates).

3. Results and discussion

3.1. Characterization of Pseudomonas putida BD2 and purification of secreted biosurfactants

The strain BD2 isolated from Arctic soil was Gram-negative, aerobic, oxidase-

positive and these results classified it as belonging to the Pseudomonas genus. The optimal

temperature for growth of strain BD2 in MSM with 2% glucose was 28 oC. Comparison of

16S rRNA nucleotide sequence from strain BD2 with sequences in the GenBank database

revealed 100% identity with the corresponding sequences from Pseudomonas putida [22].

When P. putida BD2 was grown at 28 oC in submerged culture in a sterile mineral salts

medium with 2% (w/v) glucose, 0.15 g/L of the crude biosurfactant was produced after 7 days

of culture (Fig. 1). The maximum decrease in the surface tension of the culture medium (45.8

mN/m) occurred after 48 h. The emulsification index (E24) of the culture filtrate with

petroleum ether increased from 2% to 62.9% and stabilized after 144 h. The extracted

biosurfactants were then subjected to PLC. A series of bands were observed after PLC (Fig.

2). These fractions were divided into 4 groups and scraped off from the preparative plates, and

then eluted by a chloroform–methanol mixture (2:1) and evaporated. These 4 samples,

designated A to D, were re-dissolved in 200 μL of methanol and stored for future studies.

Further assays were carried out with these 4 samples in order to determine their surface

activity by the drop-collapsing assay. Due to their strong surface tension reducing activity,

samples A (5 mg) and B (20.5 mg) from 40 mg crude extract were selected and subjected to

further analytical UPLC. Four more cycles of the biosurfactant separation and extraction

experiments had to be repeated in order to obtain sufficient quantity of samples for the

analytical UPLC, physico-chemical and biological assays.

Until recently, Pseudomonas aeruginosa was considered as potentially the main

producer of rhamnolipids [23, 24]. However, P. aeruginosa species is pathogenic. Thus, our

research was focused on characterizing other RLs producer species. This paper describes a

new strain of P. putida BD2 which is generally considered non-pathogenic [25]. Besides

production of the rhamnolipid, another important tenet is the finding that P. putida BD2 also

produces phospholipids.

Page 9: Identification and characterization of biosurfactants

Page 8 of 25

Accep

ted

Man

uscr

ipt

8

Strain BD2 grew in culture media containing different types of carbon source (data not

shown) but maximum cell growth and biosurfactant production occurred with glucose used as

a sole carbon source. While this finding was in agreement with that of Rendell et al. [26],

some other strains have shown the highest biosurfactant production on hydrophobic substrates

such as soybean oil [27, 28]. To our knowledge, non-pathogenic P. putida strains produce less

rhamnolipids than the pathogenic P. aeruginosa. The production of rhamnolipids was

observed when a strain of P. putida was grown on soluble substrates such as glucose,

molasses or on poorly soluble substrates such as hexadecane, reaching values from 0.52 to 1.2

g/L; however the exact structures of the produced rhamnolipids were not determined [29, 30].

3.2. UPLC-MS analysis

The molecular mass of the purified active compounds, i.e. fractions A and B, was

measured using UPLC/ESI-MS/MS. Previous studies revealed that the active ingredient of

biosurfactants produced by Pseudomonas was rhamnolipids [31, 32]. UPLC/MS analysis

revealed the presence of one major component with retention time 5.23 min. The mass

spectrum and chemical structure of the component is illustrated in Fig. 3A. One [M-H]-

pseudomolecular ion with m/z 649 was observed. ESI-MS showed a molecular ion of Rha-

Rha-C10-C10, which gave rise to fragments with m/z 479 and 310. The possible di-

rhamnolipid structure and fragmentation is also shown in Fig. 3A.

In the generally accepted pathway for the rhamnolipid synthesis, Rha-C10-C10 is the

precursor for Rha-Rha-C10-C10 [33]. This would suggest that for each di-rhamnolipid

detected, the mono-rhamnolipid congener should also be detected, although this is not always

the case [34]. In our work the di-rhamnolipid did not have a mono-rhamnolipid congener.

Rhamnolipids containing only one hydroxyaliphatic acid were also found by other authors

[35, 36]. In our study we did not find glycolipids with one -hydroxy fatty acid connected,

which might be due to differences either in the culture conditions or in the strain used.

ESI-MS spectra of P. putida BD2 phospholipids showed that the dominant species are

protonated phosphatidylethanolamines (PE(32:1)) (PE(33:1)) at m/z 690.5 and 704.5,

respectively. Dimers of phospholipids were observed in the range of m/z 1300–1500 and

verified by tandem mass spectrometry (MS/MS). Tandem mass spectra display the

fragmentation pattern of [PE(16:1/16:0) + H]+ and [PE(16:1/17:0) + H]+, which mainly lose

neutral phosphoethanolamine (141 mass units) to yield the fragment ion at m/z 549.5 and

563.5 respectively. The peak of m/z 549.5 was dissociated to form two major fragment ions to

Page 10: Identification and characterization of biosurfactants

Page 9 of 25

Accep

ted

Man

uscr

ipt

9

form [C16H31O]+ of m/z 239.2, and [C16H29O]+ of m/z 237.2 (Fig. 3B). The m/z 563.5 was

used as precursor ions for further ESI-MS/MS analyses (Fig. 3C).

3.3. TLC analysis

Thin-layer chromatography showed that the crude extract was a mixture of compounds

that could be classified as glycolipid and phospholipid. Visualization was carried out with

orcinol and ninhydrin followed by heating. The lower spot (Rf = 0.26) is from the di-

rhamnolipid, while the higher spot (Rf = 0.47) comes from phosphatidylethanolamines (data

not shown). Our results are similar to previously reported results of TLC of rhamnolipids

from different strains of Pseudomonas [23]. The Rf values obtained by these authors are 0.27

and 0.57 for di- and mono-rhamnolipid, respectively.

3.4. Physico-chemical properties of biosurfactants

Biosurfactants tend to form aggregates with their hydrophilic head groups in contact

with water and their hydrophobic tails directed toward the interior of the micelles and away

from water. These aggregates are generally called micelles. Depending on their shape and

structure, micelles can be generally divided into four classes: spheres, rods, lamellar

structures, and vesicles. The aggregation behavior of Rha-Rha-C10-C10 was characterized by

measuring the change in the size of aggregates with changing compound concentration. Fig. 4

shows the aggregation behavior of Rha-Rha-C10-C10. The size of the aggregates increased

with concentration in the range of 0.15–0.5 mg/mL and reached 80-121 nm. In previous

studies, we have demonstrated that aggregation above the CMC was also observed for

pseudofactin II, a surface active lipopeptide produced by Pseudomonas fluorescens [37]. The

mean hydrodynamic diameter of pseudofactin-biosurfactant micelles in water ranged from

40.2 nm to 60.3 nm. Xu et al. [38] demonstrated that the type and concentration of surfactant

significantly influence the properties of aggregates. The aggregation behavior of the surface

active compounds affects the size distribution and stability of aggregates. There is still little

information available about aggregation and polydispersity index (PDI) of various

rhamnolipids. In this work, we show the size of aggregates for one type of rhamnolipid (Rha-

Rha-C10-C10), while other reports focused on mixtures of rhamnolipid homologs [39, 40].

The new data about each described biosurfactant will contribute to understanding the

aggregation process and compounds the final activity of the product.

Page 11: Identification and characterization of biosurfactants

Page 10 of 25

Accep

ted

Man

uscr

ipt

10

The emulsifying properties of 0.2% aqueous solution of biosurfactants produced by

BD2 strain were examined with different vegetable oils and hydrocarbons. The purified

phospholipids and the rhamnolipid emulsified vegetable oils more efficiently than

hydrocarbons (Fig. 5). Biosurfactants could emulsify 70% of olive oil, while emulsification of

n-hexane, xylene, hexadecane, and petroleum ether ranged from 51 to 65%.

The rhamnolipid (Rha-Rha C10 C10) reduced surface tension of the water from 72 to

31 mN/m. Critical micelle concentration (CMC) determined with the aid of a series of

concentrations was around 0.130 mg/mL (Fig. 6). The surface tension properties of

rhamnolipids are dependent on bacterial strain, medium composition, and culture conditions

that determine the composition and distribution of homolog molecules present in the final

product [41]. CMC values in a wide range from 5 to 386 mg/L and surface tension from 25 to

31 mN/m have been reported for different rhamnolipid mixtures [42]. Abalos et al. [43]

reported CMC values of 106, 150 and 234 mg/L for different mixtures of mono- and di-

rhamnolipids. More hydrophilic rhamnolipids like Rha-C10 or Rha-Rha-C10 yielded CMC as

high as 200 mg/L whereas lower values of 5-60 mg/L have been reported for mixtures

containing mainly mono-rhamnolipid Rha-C10-C10 [44].

3.5. Biological activity of biosurfactants

The microtiter-plate based anti-adhesion assay estimated the rhamnolipid and the

phospholipids concentrations that were effective in inhibiting adhesion of the pathogenic

microorganisms. The highest percentage of microbial adhesion inhibition was obtained

against C. albicans SC5314, 90% for the rhamnolipid (Table 1) and 98% for the

phospholipids (Table 2) at 0.5 mg/mL concentration of the biosurfactants. The order of anti-

adhesive action against microorganisms studied was C. albicans>E. faecalis>P. mirabilis>E.

coli>S. epidermidis>E. hirae.

Attempts to reduce or inhibit microbial adherence is a viable means to control

infection, since such adherence is one of the initial stages of the infectious process. Thus, pre-

coating of solid surfaces by biosurfactants might constitute a new and effective means of

combating colonization by pathogens and an effective strategy to reduce microbial adhesion

[10]. Thanks to their amphiphilic structure, surface active compounds reduce the surface

tension at interfaces, and thus affect the adhesion and detachment of microorganisms [10].

Gudina et al. [45] characterized the anti-adhesive activity of biosurfactants against several

microorganisms including Gram-positive and Gram-negative bacteria. Several investigations

have pointed that biosurfactants produced by probiotic microorganisms inhibited adhesion of

Page 12: Identification and characterization of biosurfactants

Page 11 of 25

Accep

ted

Man

uscr

ipt

11

bacteria or fungi to artificial surfaces [46-48]. For example, biosurfactants from Streptococcus

thermophilus and Lactobacillus acidophilus suppressed C. albicans adhesion to silicone

rubber [8, 49].

In conclusion, we suggested that the phospholipids and rhamnolipid produced by

nonpathogenic P. putida BD2 could be used as alternative anti-adhesive agents against

pathogenic microorganisms responsible for diseases and infections in the urinary, vaginal and

gastrointestinal tracts. This points to their potential utility as coating agents for medical

insertional materials.

4. Conclusions

Our data allow several conclusions:

- The new bacterial isolate BD2 was identified as Pseudomonas putida.

- Two distinct fractions of the crude biosurfactants correspond to a mixture of

phosphatidylethanolamines PE(32:1), PE(33:1) and di-rhamnolipid (Rha-Rha–C10–C10).

- P. putida species has not been previously described as rhamnolipid and phospholipid

producer.

- The physico-chemical properties such as CMC (0.130 mg/mL), surface tension (31 mN/m),

emulsification (51- 70%), and size of aggregates above CMC 80-121 nm were evaluate.

- Described biosurfactants have strong anti-adhesive activities against bacterial and yeast

strains on a polystyrene surface.

Acknowledgments

This work was supported by grant from the University of Wroclaw 1498/M/WB/11 and

by the Polish National Centre for Science No N N302 640940.

Page 13: Identification and characterization of biosurfactants

Page 12 of 25

Accep

ted

Man

uscr

ipt

12

References

[1] J.D. Desai and I.M. Banat, Microbial production of surfactants and their commercial potential, Microbiol Mol Biol Rev, 61 (1997) 47-64.[2] N. Christofi and I.B. Ivshina, Microbial surfactants and their use in field studies of soil remediation, J Appl Microbiol, 93 (2002) 915-929.[3] E.Z. Ron and E. Rosenberg, Natural roles of biosurfactants, Environ Microbiol, 3 (2001) 229-236.[4] J.C. Mata-Sandoval, J. Karns and A. Torrents, High-performance liquid chromatography method for the characterization of rhamnolipid mixtures produced by pseudomonas aeruginosa UG2 on corn oil, J Chromatogr A, 864 (1999) 211-220.[5] I.M. Banat, R.S. Makkar and S.S. Cameotra, Potential commercial applications of microbial surfactants, Appl Microbiol Biotechnol, 53 (2000) 495-508.[6] G. Georgiou, S.C. Lin and M.M. Sharma, Surface-active compounds from microorganisms, Biotechnology (N Y), 10 (1992) 60-65.[7] T. Janek, M. Lukaszewicz and A. Krasowska, Antiadhesive activity of the biosurfactant pseudofactin II secreted by the Arctic bacterium Pseudomonas fluorescens BD5, BMC Microbiol, 12 (2012) 24.[8] H.J. Busscher, C.G. van Hoogmoed, G.I. Geertsema-Doornbusch, M. van der Kuijl-Booij and H.C. van der Mei, Streptococcus thermophilus and its biosurfactants inhibit adhesion by Candida spp. on silicone rubber, Appl Environ Microbiol, 63 (1997) 3810-3817.[9] G. Reid, J.A. Younes, H.C. Van der Mei, G.B. Gloor, R. Knight and H.J. Busscher, Microbiota restoration: natural and supplemented recovery of human microbial communities, Nat Rev Microbiol, 9 (2011) 27-38.[10] L. Rodrigues, I.M. Banat, J. Teixeira and R. Oliveira, Biosurfactants: potential applications in medicine, J Antimicrob Chemother, 57 (2006) 609-618.[11] G. Soberon-Chavez, F. Lepine and E. Deziel, Production of rhamnolipids by Pseudomonas aeruginosa, Appl Microbiol Biotechnol, 68 (2005) 718-725.[12] M. Nitschke, S.G. Costa and J. Contiero, Rhamnolipid surfactants: an update on the general aspects of these remarkable biomolecules, Biotechnol Prog, 21 (2005) 1593-1600.[13] M.M. Muller, J.H. Kugler, M. Henkel, M. Gerlitzki, B. Hormann, M. Pohnlein, C. Syldatk and R. Hausmann, Rhamnolipids--next generation surfactants?, J Biotechnol, 162 (2012) 366-380.[14] L. Sim, O.P. Ward and Z.Y. Li, Production and characterisation of a biosurfactant isolated from Pseudomonas aeruginosa UW-1, J Ind Microbiol Biotechnol, 19 (1997) 232-238.[15] M. Henkel, M.M. Müller, J.H. Kügler, R.B. Lovaglio, J. Contiero, C. Syldatk and R. Hausmann, Rhamnolipids as biosurfactants from renewable resources: Concepts for next-generation rhamnolipid production., Process Biochemistry, 47 (2012) 1207-1219.[16] M. Partovi, T.B. Lotfabad, R. Roostaazad, M. Bahmaei and S. Tayyebi, Management of soybean oil refinery wastes through recycling them for producing biosurfactant using Pseudomonas aeruginosa MR01, World J Microbiol Biotechnol.[17] J.Y. Wu, K.L. Yeh, W.B. Lu, C.L. Lin and J.S. Chang, Rhamnolipid production with indigenous Pseudomonas aeruginosa EM1 isolated from oil-contaminated site, Bioresour Technol, 99 (2008) 1157-1164.[18] D.G. Cooper and B.G. Goldenberg, Surface-active agents from two bacillus species, Appl Environ Microbiol, 53 (1987) 224-229.[19] D.K. Jain, D.L. Collins-Thompson, H. Lee and J.T. Trevors, A drop-collapsing test for screening surfactant-producing microorganisms, J Microbiol Meth, 13 (1991) 271-279.[20] T. Janek, M. Lukaszewicz, T. Rezanka and A. Krasowska, Isolation and characterization of two new lipopeptide biosurfactants produced by Pseudomonas fluorescens BD5 isolated

Page 14: Identification and characterization of biosurfactants

Page 13 of 25

Accep

ted

Man

uscr

ipt

13

from water from the Arctic Archipelago of Svalbard, Bioresour Technol, 101 (2010) 6118-6123.[21] Y.H. Hsueh, E.B. Somers, D. Lereclus, E. Ghelardi and A.C. Wong, Biosurfactant production and surface translocation are regulated by PlcR in Bacillus cereus ATCC 14579 under low-nutrient conditions, Appl Environ Microbiol, 73 (2007) 7225-7231.[22] H. Tang, Y. Yao, L. Wang, H. Yu, Y. Ren, G. Wu and P. Xu, Genomic analysis of Pseudomonas putida: genes in a genome island are crucial for nicotine degradation, Sci Rep, 2 (2012) 377.[23] H. Abbasi, M.M. Hamedi, T.B. Lotfabad, H.S. Zahiri, H. Sharafi, F. Masoomi, A.A. Moosavi-Movahedi, A. Ortiz, M. Amanlou and K.A. Noghabi, Biosurfactant-producing bacterium, Pseudomonas aeruginosa MA01 isolated from spoiled apples: physicochemical and structural characteristics of isolated biosurfactant, J Biosci Bioeng, 113 (2011) 211-219.[24] A.M. Abdel-Mawgoud, F. Lepine and E. Deziel, Rhamnolipids: diversity of structures, microbial origins and roles, Appl Microbiol Biotechnol, 86 (2010) 1323-1336.[25] E. Anaissie, V. Fainstein, P. Miller, H. Kassamali, S. Pitlik, G.P. Bodey and K. Rolston, Pseudomonas putida. Newly recognized pathogen in patients with cancer, Am J Med, 82 (1987) 1191-1194.[26] N.B. Rendell, G.W. Taylor, M. Somerville, H. Todd, R. Wilson and P.J. Cole, Characterisation of Pseudomonas rhamnolipids, Biochim Biophys Acta, 1045 (1990) 189-193.[27] H. Abbasi, K.A. Noghabi, M.M. Hamedi, H.S. Zahiri, A.A. Moosavi-Movahedi, M. Amanlou, J.A. Teruel and A. Ortiz, Physicochemical characterization of a monorhamnolipid secreted by Pseudomonas aeruginosa MA01 in aqueous media. An experimental and molecular dynamics study, Colloids Surf B Biointerfaces, 101 (2012) 256-265.[28] Y. Lee, S.Y. Lee and J.W. Yang, Production of rhamnolipid biosurfactant by fed-batch culture of Pseudomonas aeruginosa using glucose as a sole carbon source, Biosci Biotechnol Biochem, 63 (1999) 946-947.[29] B.K. Tuleva, G.R. Ivanov and N.E. Christova, Biosurfactant production by a new Pseudomonas putida strain, Z Naturforsch C, 57 (2002) 356-360.[30] D. Onbasli and B. Aslim, Biosurfactant production in sugar beet molasses by somePseudomonas spp, J Environ Biol, 30 (2009) 161-163.[31] A.A. Bodour, C. Guerrero-Barajas, B.V. Jiorle, M.E. Malcomson, A.K. Paull, A. Somogyi, L.N. Trinh, R.B. Bates and R.M. Maier, Structure and characterization of flavolipids, a novel class of biosurfactants produced by Flavobacterium sp. strain MTN11, Appl Environ Microbiol, 70 (2004) 114-120.[32] T.B. Lotfabad, H. Abassi, R. Ahmadkhaniha, R. Roostaazad, F. Masoomi, H.S. Zahiri, G. Ahmadian, H. Vali and K.A. Noghabi, Structural characterization of a rhamnolipid-type biosurfactant produced by Pseudomonas aeruginosa MR01: enhancement of di-rhamnolipid proportion using gamma irradiation, Colloids Surf B Biointerfaces, 81 (2010) 397-405.[33] M.M. Burger, L. Glaser and R.M. Burton, The Enzymatic Synthesis of a Rhamnose-Containing Glycolipid by Extracts of Pseudomonas Aeruginosa, J Biol Chem, 238 (1963) 2595-2602.[34] A. Manso Pajarron, C.G. De Koster, W. Heerma, Schmidt, M. and J. Haverkamp, Structure identification of natural rhamnolipid mixtures by fast-atom-bombardment tandem massspectrometry, Glycoconjug. J., 10 (1993) 219–226.[35] E. Haba, A. Pinazo, O. Jauregui, M.J. Espuny, M.R. Infante and A. Manresa, Physicochemical characterization and antimicrobial properties of rhamnolipids produced by Pseudomonas aeruginosa 47T2 NCBIM 40044, Biotechnol Bioeng, 81 (2003) 316-322.[36] E. Deziel, F. Lepine, D. Dennie, D. Boismenu, O.A. Mamer and R. Villemur, Liquid chromatography/mass spectrometry analysis of mixtures of rhamnolipids produced by

Page 15: Identification and characterization of biosurfactants

Page 14 of 25

Accep

ted

Man

uscr

ipt

14

Pseudomonas aeruginosa strain 57RP grown on mannitol or naphthalene, Biochim Biophys Acta, 1440 (1999) 244-252.[37] T. Janek, A. Krasowska, A. Radwanska and M. Lukaszewicz, Lipopeptide Biosurfactant Pseudofactin II Induced Apoptosis of Melanoma A 375 Cells by Specific Interaction with the Plasma Membrane, PLoS One, 8 (2013) e57991.[38] Q. Xu, M. Nakajima, S. Ichikawa, N. Nakamura, P. Roy, H. Okadome and T. Shiina, Effects of surfactant and electrolyte concentrations on bubble formation and stabilization, J Colloid Interface Sci, 332 (2009) 208-214.[39] M. Sanchez, F.J. Aranda, M.J. Espuny, A. Marques, J.A. Teruel, A. Manresa and A. Ortiz, Aggregation behaviour of a dirhamnolipid biosurfactant secreted by Pseudomonas aeruginosa in aqueous media, J Colloid Interface Sci, 307 (2007) 246-253.[40] Y.P. Guo, Y.Y. Hu, R.R. Gu and H. Lin, Characterization and micellization of rhamnolipidic fractions and crude extracts produced by Pseudomonas aeruginosa mutant MIG-N146, J Colloid Interface Sci, 331 (2009) 356-363.[41] M. Nitschke, S.G. Costa and J. Contiero, Structure and applications of a rhamnolipid surfactant produced in soybean oil waste, Appl Biochem Biotechnol, 160 (2009) 2066-2074.[42] S. Lang and D. Wullbrandt, Rhamnose lipids--biosynthesis, microbial production and application potential, Appl Microbiol Biotechnol, 51 (1999) 22-32.[43] A. Abalos, A. Pinazo, M.R. Infante, M. Casals, F. Garcı´a and A. Manresa, Physicochemical and antimicrobial properties of new rhamnolipids produced by Pseudomonas aeruginosa AT10 from soybean oil refinery wastes., Langmuir, 17 (2001) 1367–1371.[44] M.I. Van Dyke, P. Couture, M. Brauer, H. Lee and J.T. Trevors, Pseudomonas aeruginosa UG2 rhamnolipid biosurfactants: structural characterization and their use in removing hydrophobic compounds from soil, Can J Microbiol, 39 (1993) 1071-1078.[45] E.J. Gudina, J.A. Teixeira and L.R. Rodrigues, Isolation and functional characterization of a biosurfactant produced by Lactobacillus paracasei, Colloids Surf B Biointerfaces, 76 (2010) 298-304.[46] M.M. Velraeds, B. van de Belt-Gritter, H.J. Busscher, G. Reid and H.C. van der Mei, Inhibition of uropathogenic biofilm growth on silicone rubber in human urine by lactobacilli--a teleologic approach, World J Urol, 18 (2000) 422-426.[47] M.M. Velraeds, H.C. van der Mei, G. Reid and H.J. Busscher, Inhibition of initial adhesion of uropathogenic Enterococcus faecalis to solid substrata by an adsorbed biosurfactant layer from Lactobacillus acidophilus, Urology, 49 (1997) 790-794.[48] M.M. Velraeds, H.C. van der Mei, G. Reid and H.J. Busscher, Inhibition of initial adhesion of uropathogenic Enterococcus faecalis by biosurfactants from Lactobacillus isolates, Appl Environ Microbiol, 62 (1996) 1958-1963.[49] M.M. Velraeds, B. van de Belt-Gritter, H.C. van der Mei, G. Reid and H.J. Busscher, Interference in initial adhesion of uropathogenic bacteria and yeasts to silicone rubber by a Lactobacillus acidophilus biosurfactant, J Med Microbiol, 47 (1998) 1081-1085.

Page 16: Identification and characterization of biosurfactants

Page 15 of 25

Accep

ted

Man

uscr

ipt

15

Tables

Table 1. Inhibition of microbial adhesion in the microtiter plate by purified di-rhamnolipid.

PBS was used as control and set at 0% as no inhibition occurs. Values ± confidence interval,

n=9, =0.05.

Microorganism Inhibition of microbial adhesion (%)

Rhamnolipid concentration (mg/ml) Control (PBS)

0.500 0.250 0.150 0.075 0.035 0

Escherichia coliATCC 25922 65±0.2 30±0.2 21±0.1 13±0.2 9±0.1 0

Escherichia coliATCC 10536 52±0.2 44±0.1 42±0.1 42±0.1 39±0.2 0

Escherichia coli17-2 62±0.1 61±0.2 58±0.5 41±0.1 34±0.3 0

Enterococcus faecalis ATCC 29212 79±0.3 67±0.1 55±0.1 10±0.2 7±0.2 0

Enterococcus faecalis JA/3 55±0.1 26±0.1 18±0.2 16±0.3 8±0.2 0

Enterococcus hiraeATCC 10541 43±0.1 21±0.1 19±0.3 14±0.3 9±0.3 0

Staphylococcus epidermidis KCTC 1917 45±0.1 41±0.1 22±0.1 11±0.1 10±0.2 0

Proteus mirabilisATCC 21100 75±0.5 72±0.1 71±0.1 65±0.1 54±0.1 0

Candida albicans ATCC 20231 89±0.2 84±0.2 72±0.1 36±0.1 25±0.1 0

Candida albicansSC5314 90±0.2 72±0.2 69±0.1 40±0.1 37±0.2 0

Page 17: Identification and characterization of biosurfactants

Page 16 of 25

Accep

ted

Man

uscr

ipt

16

Table 2. Inhibition of microbial adhesion in the microtiter plate by purified

phosphatidylethanolamines. PBS was used as control and set at 0% as no inhibition occurs.

Values ± confidence interval, n=9, =0.05.

Microorganism Inhibition of microbial adhesion (%)

Phosphatidylethanolamines concentration (mg/ml) Control (PBS)

0.500 0.250 0.150 0.075 0.035 0

Escherichia coliATCC 25922 35±0.2 20±0.1 11±0.1 9±0.2 4±0.2 0

Escherichia coliATCC 10536 47±0.2 42±0.2 38±0.2 24±0.1 19±0.2 0

Escherichia coli17-2 42±0.2 36±0.2 28±0.1 25±0.2 19±0.1 0

Enterococcus faecalis ATCC 29212 72±0.1 70±0.1 54±0.1 43±0.1 38±0.3 0

Enterococcus faecalis JA/3 40±0.2 31±0.1 25±0.3 9±0.2 2±0.3 0

Enterococcus hiraeATCC 10541 23±0.1 19±0.2 15±0.1 11±0.3 8±0.1 0

Staphylococcus epidermidis KCTC 1917 41±0.2 38±0.1 33±0.2 14±0.1 6±0.2 0

Proteus mirabilisATCC 21100 63±0.2 58±0.1 52±0.1 45±0.2 35±0.2 0

Candida albicans ATCC 20231 96±0.1 88±0.2 82±0.2 76±0.1 75±0.1 0

Candida albicansSC5314 98±0.1 93±0.1 90±0.1 84±0.1 79±0.1 0

Page 18: Identification and characterization of biosurfactants

Page 17 of 25

Accep

ted

Man

uscr

ipt

17

Figures

Fig. 1. Time course of biosurfactant production, cell growth, surface tension and

emulsification activity of P. fluorescens BD5 grown on mineral salt medium with 2% glucose

as carbon source at 28 oC.

Fig. 2. Preparative layer chromatography (PLC) of compounds obtained from P. putida BD2

by using a mobile phase of chloroform/methanol/water in the 65:15:2 ratio (vol/vol/vol).

Fractions were visualized by UV transilluminator.

Fig. 3. UPLC-ESI–MS spectrum, negative and positive ion mode of rhamnolipid and

phospholipids respectively. (a) Negative mass spectrometry showing the predominance

(retention time of 5.23 min) of m/z 649 (Rha-Rha-C10-C10). Product ion spectra of the

protonated molecules [M + H]+ of (b) PE(32:1) at retention time 10.13 min and (c) PE(33:1)

at retention time 10.64 min.

Fig. 4. The effect of di-rhamnolipid concentration on aggregate size.

Fig. 5. Emulsifying index (E24) of biosurfactants with different hydrocarbons, mean ± SD

(n=3).

Fig. 6. Effect of Rha-Rha-C10-C10 concentration on surface tension. The CMC was

determined from the intersection of regression lines that describe the two parts of the curve,

below and above CMC. Results represent the average of three independent measurements.

Page 19: Identification and characterization of biosurfactants

Page 18 of 25

Accep

ted

Man

uscr

ipt

18

Graphical abstract

Highlights

1. Pseudomonas putida BD2 was isolated and identified from an Arctic soil.

2. The strain was found to release to growth medium biosurfactants.

3. Biosurfactants were purified and indentified using various physico-chemical methods.

4. Structures were identified using UPLC-MS method.

5. One glycolipid and two phospholipids were identified.

Page 20: Identification and characterization of biosurfactants

Page 19 of 25

Accep

ted

Man

uscr

ipt

*Graphical Abstract (for review)

Page 21: Identification and characterization of biosurfactants

Page 20 of 25

Accep

ted

Man

uscr

ipt

2

Fig. 1.

Page 22: Identification and characterization of biosurfactants

Page 21 of 25

Accep

ted

Man

uscr

ipt

3

Fig. 2.

Page 23: Identification and characterization of biosurfactants

Page 22 of 25

Accep

ted

Man

uscr

ipt

4

Fig. 3.

Fig. 4.

Page 24: Identification and characterization of biosurfactants

Page 23 of 25

Accep

ted

Man

uscr

ipt

5

Page 25: Identification and characterization of biosurfactants

Page 24 of 25

Accep

ted

Man

uscr

ipt

6

Fig. 5.

Page 26: Identification and characterization of biosurfactants

Page 25 of 25

Accep

ted

Man

uscr

ipt

7

Fig. 6.