Integrated Metabolomic and Proteomic Analysis Reveals Systemic

Nov 12, 2014 - Department of Plant Sciences, School of Life Sciences, University of Hyderabad, P.O. Central University, Hyderabad 500 046, India. ‡...
0 downloads 0 Views 5MB Size
Subscriber access provided by Library, Special Collections and Museums, University of Aberdeen

Article

Integrated metabolomic and proteomic analysis reveals systemic responses of Rubrivivax benzoatilyticus JA2 to aniline stress Md Mujahid, M Lakshmi Prasuna, Ch Sasikala, and Ch Venkata Ramana J. Proteome Res., Just Accepted Manuscript • DOI: 10.1021/pr500725b • Publication Date (Web): 12 Nov 2014 Downloaded from http://pubs.acs.org on November 16, 2014

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Journal of Proteome Research is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

Integrated metabolomic and proteomic analysis reveals systemic responses of Rubrivivax benzoatilyticus JA2 to aniline stress Md Mujahid§ , M Lakshmi Prasuna§, Ch Sasikala±, Ch Venkata Ramana§*

§

Department of Plant Sciences, School of Life Sciences, University of Hyderabad, P.O.

Central University, Hyderabad 500 046, India. ±

Bacterial Discovery Laboratory, Center for Environment, IST, JNT University Hyderabad,

Kukatpally, Hyderabad 500 085, India. *

Corresponding author:

Prof. Ch. V. Ramana, Department of Plant Sciences, School of Life Sciences, University of Hyderabad. Hyderabad-500 046, Telangana, India. E-mail: [email protected]; [email protected] Tel phone : +91 040 23134502 Fax: +91 040 23010120 & 23010145,

Running title: Systemic responses of Rubrivivax benzoatilyticus JA2 to aniline

1

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 54

Abstract Aromatic amines are widely distributed in the environment and are major environmental pollutants. Although degradation of aromatic amines is well studied in bacteria, physiological adaptations and stress response to these toxic compounds is not yet fully understood. In the present study, systemic responses of Rubrivivax benzoatilyticus JA2 to aniline stress were deciphered using metabolite and iTRAQ-labelled protein profiling. Strain JA2 tolerated high concentrations of aniline (30 mM) with trace amounts of aniline being transformed to acetanilide. GC-MS metabolite profiling revealed aniline stress phenotype wherein amino acid, carbohydrate, fatty acid, , nitrogen metabolisms and TCA (tricarboxylic acid cycle) were modulated. Strain JA2 responded to aniline by remodelling the proteome and

cellular functions such as signalling, transcription, translation, stress

tolerance, transport and carbohydrate metabolism were highly modulated. Key adaptive responses such as transcription/translational changes, molecular chaperones to control protein folding and efflux pumps implicated in solvent extrusion were induced in response to aniline stress. Proteo-metabolomics indicated extensive re-wiring of metabolism to aniline. TCA cycle and amino acid catabolism were down-regulated while gluconeogenesis and pentose phosphate pathways were up-regulated leading to the synthesis of extracellular polymeric substances. Furthermore, increased saturated fatty acid ratios in membrane due to aniline stress suggest membrane adaptation. The present study thus indicates that

strain JA2

employs multi-layered responses; stress response, toxic compound tolerance, energy conservation and metabolic rearrangements to aniline. Key words. GC-MS, iTRAQ, Rubrivivax benzoatilyticus JA2, aniline stress, systemtic responses, Extracellular polymeric substances.

2

ACS Paragon Plus Environment

Page 3 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

Introduction Aromatic amines represent one of the most important classes of anthropogenic compounds and are common by-products of pesticides, dyestuffs, rubbers or pharmaceutical manufacturing in addition to coal and fossil fuel combustion.1, 2 These are widely distributed in the environment due to their extensive usage in various industrial and domestic applications and highly persistent due to their unique chemical composition.1,3

Many

aromatic amines are deleterious to life forms due to their genotoxic and/or cytotoxic properties1, 4 and accounts for 12% of chemicals that are carcinogens.2, 4 One such aromatic amine is aniline, which is extensively used in manufacturing of dyestuffs, plastics, rubber, herbicides, pesticides, paints and pharmaceuticals2, 5 and it is continuously released into the environment through industrial effluents as well as during use. Owing to its extensive usage aniline is widely distributed and often found in both terrestrial and aquatic environments.6, 7 Aniline is a recalcitrant xenobiotic compound; is listed as priority pollutant and its fate in the environment has been a serious concern as it is toxic to life forms.5,8 Many aniline degrading5, 8

or transforming bacteria were isolated 9 and their biodegradation pathways under both oxic5,

7

and anoxic7, 10 conditions are well reported. Though aniline acts as a potential carbon source

to bacteria, it acts as a stressor to both degraders and non-degraders and bacterial stress responses to aniline are not yet elucidated. Cellular adaptations/tolerance mechanisms to xenobiotics enable bacteria to thrive in the presence of toxic compounds and degrade them.11, 12

Thus gaining insights into toxic compound tolerance may eventually lead to the

development of effective bioremediation processes.13 Several studies on model organisms have attempted to explain the mechanism of toxic compound tolerance using genomic,11, 14, 15 transcriptomic16,

17

and proteomic18-20 approaches. Though informative, these approaches

identified only a small subset of predicted proteins in model organisms and these studies are largely confined to xenobiotics like toluene,18,

19

ethylbenzene,18 xylene,16 benzoate,16

benzene,20 polychlorinated biphenyl,21 cyclohexane.22 3

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 54

However, how bacteria cope with aniline, and their possible stress response or tolerance mechanisms are not elucidated so far. Previous studies reported that an anoxygenic photosynthetic bacterium, Rubrivivax benzoatilyticus JA2 grows in the presence of high concentrations of aniline and other aromatic compounds.23 Although R. benzoatilyticus JA2 could not utilize aniline as a sole source of carbon or nitrogen, it tolerated high concentrations of aniline (20-30 mM) and its mono, di-substituted derivatives.24 Tolerance to

high

concentrations of anilines without metabolizing it is an intriguing aspect and in the present study, we used R. benzoatilyticus JA2 as a model organism to elucidate the tolerance/stress response mechanisms to aniline. GC-MS based global metabolite profiling followed by multivariate analysis revealed metabolic responses of strain JA2 in the form of key metabolites (presence and quantities) altered in the presence of aniline. Further, iTRAQ based quantitative proteomics provided the insights of molecular adaptations to aniline. Finally, integrated metabolomic and proteomic studies revealed metabolic shift or metabolic reprogramming of cell to aniline stress towards cell survival. Our data also suggest that general stress response and multiple tolerance strategies for cell survival were evoked under aniline stress. Materials and methods Organism and growth conditions Rubrivivax benzoatilyticus JA2 (ATCC

BBA-35)

was grown photoheterotrophically

(anaerobic, 30 ºC; light 2,400 lux) in a mineral medium25 supplemented with malate (22 mM) as carbon source and ammonium chloride (7 mM) as nitrogen source in fully filled screw cap test tubes (10x100 mm)/reagent bottles (250 ml) at pH 6.8 and a temperature of 30 ±1oC. Photoheterotrophically grown late log phase, 48 h culture (OD660nm 0.4) of strain JA2 was exposed to sterilized aniline (25 mM) and culture incubated under photoheterotrophic conditions. Cells were harvested at 4 oC and immediately frozen in liquid nitrogen and stored

4

ACS Paragon Plus Environment

Page 5 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

at -80 oC until further use. To evaluate the effect of aniline on growth, R. benzoatilyticus JA2 was grown on malate mineral medium under photoheterotrophic conditions with different concentrations of aniline (0-35 mM). Growth was monitored by measuring OD at 660 nm. Stable isotope feeding and extraction of metabolites To identify the transformation products of aniline, photoheterotrophically grown culture (0.4 OD at 660 nm) was exposed to 25 mM of unlabelled or deuterium labelled (aniline-2,3,4,5,6-d5, 98%-D, Sigma-Aldrich) aniline. After 48 h of incubation under photoheterotrophic conditions, cells were harvested by centrifugation (10,000 x g, 4 oC, 10 min), culture supernatant collected and acidified to pH 2.5 with 5N HCl. Acidified supernatant was extracted twice with equal volumes of ethyl acetate and ethyl acetate layers were pooled, evaporated to dryness under vacuum at 45 oC using flash evaporator (Heidolph, Germany). Finally, the dried residue was dissolved in HPLC grade methanol, filtered (0.2 µm membrane, Supro PALL) and stored at -20 oC for HPLC and mass analysis. HPLC and liquid chromatography mass spectrometry (LC-MS-ESI) HPLC and LC-MS analysis was done according to Mujahid et al.24 In brief, HPLC was performed on Shimadzu’s Prominence HPLC (LC-20AT) equipped with photodiode array detector. Metabolites separated by using Phenomenex reverse phase column (Luna C18, 5 µm, 250 x 4.6 mm) with a mobile phase which consisted of 1% acetic acid (solvent A) and acetonitrile (solvent B). Metabolites were eluted using linear gradient (0-100%) of acetonitrile for 30 min, at a flow rate 1.5 ml/min and metabolites detected at 260 nm. Mass spectral analysis was done on microTOF-Q (Brukers Deltonics) mass spectrometer linked to Agilent HPLC (1200 series). Metabolites were separated on C-18 column (Waters C-18, 5 µm, 150 x 4.6 mm) and separation conditions were same as described in HPLC analysis except that the flow rate was 0.8 ml/min. Electro spray ionization (ESI) positive (+) ion mode

5

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 54

was used to detect the molecular ion mass (M+H) and fragmentation obtained by using collision energy of 10 eV and mass spectra recorded at 50-500 Da. Extraction of intracellular metabolites Aniline exposed and control cells (three independent experiments, each with two biological replicates, a total of 6 samples used per condition) were harvested by centrifugation (10,000 x g, 10 min, 4 ○C) and the cell pellet was washed with cold Milli-Q water, rapidly quenched in liquid nitrogen, and the cell pellet was lyophilized (Freeze Dryer, Labcanco USA) for 10 h. Metabolites were extracted from the lyophilized sample by methanol/chloroform method.26 Nine millilitres of methanol:chloroform:water (3:1:1 v/v/v) mixture was added to lyophilized cell pellet vortexed and sonicated for 15 sec, 7 times (8 cycle, 50% power, 4°C) with time interval of 1 min. Chloroform (1.5 ml) and 3 ml of water were added to sonicated sample, vortexed and centrifuged (12,000 x g, 4°C, 10 mins). A final polar phase which contains hydrophilic metabolites were collected, methanol was evaporated from the sample and lyophilized and stored at -20°C until further analysis by GC-MS. Sample derivatization and GC-MS analysis Metabolites were derivatized according to Jozefczuk et a.,27 Samples were first derivatized by adding 20 µl of 40 mg.ml-1 methoxyamine hydrochloride (Sigma-Aldrich) in pyridine (Sigma-Aldrich) and incubated at 30°C for 90 min, followed by adding 40 µl of BSTFA (N,O-Bis(trimethylsilyl)trifluoroacetamide) and TMCS (99:1) and incubated at 37°C for 30 min. GC-MS analysis was performed on Pegasus HT TOF-MS (Leco, USA) system equipped with Agilent series (7890) gas chromatography. 1 µl of derivatized sample was injected into HP-5 column (30 m, internal diameter 0.32 mm, thickness 0.25 µm), with helium as carrier gas at a constant flow of 1.2 ml.min-1 in split less mode. Initial oven temperature was held at 80°C for 2 min, ramped to 180°C by 3°C min-1 held for 1 min finally ramped to 310○C by 4○C min-1 and isocratic hold for 3 min (310○C). Inlet temperature was 6

ACS Paragon Plus Environment

Page 7 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

250○C, transfer line temperature 225°C, source temperature 230°C and ionization energy -70 eV. Mass spectra were recorded at 40-1000 m/z with 3 spectra /sec. Chromatograms were processed using Leco ChromaTOF software (version 4.21). A reference chromatogram was defined as that which had a maximum of detected peaks over a signal/noise ratio of 10 and this was used for automated peak identification. Metabolites were identified based on mass spectral comparison to the standard NIST 98 library, mass spectral matching was manually supervised and matches accepted with thresholds of match >800 (with a maximum match equal to 1000). Metabolite identity was further confirmed by comparing with mass spectral libraries of Golm Metabolome database (www.gmd.mpimpgolm.mpg.de), mass bank (www.massbank.jp) and some authentic (amino acids, carbohydrates, organic acid) standards run under identical conditions. Authentic standards mix (quality control sample) was run routinely before and along with the batch of test samples4. Technical consistency was assessed using coefficient of variations (CV) between quality control samples which was < 15%. GC-MS data processing and statistical analysis GC-MS data was further processed and subjected to statistical analysis. Relative metabolite abundances were calculated from peak areas (unique mass) of identified metabolites obtained from GC-MS analysis. Peak areas were normalized to dry weight of the sample and more than 50% of missing values excluded from the data. Data was log transformed and subjected to quantile normalization using MetaboAnalyst28 and normalized data was used for multivariate statistical analysis. Principal component analysis (PCA), partial least-squares discrimination analysis (PLS-DA) and hierarchical clustering analysis (HCA) were performed using MetaboAnalyst (www.metaboanalyst.ca/MetaboAnalyst). HCA was done using Pearson correlation as distance matrix. Data was subjected to t test analysis to identify metabolites significantly regulated between control and aniline exposed cells. Metabolites 7

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 54

having fold change >2 and P- value ≤0.05 were considered as statistically significant and metabolites were annotated to metabolic pathways according to KEGG (Kyoto Encyclopedia of Genes and Genomes, www.genome.jp/kegg) database to identify metabolic pathways influenced by aniline. Scanning electron microscopy (SEM) Scanning electron microscopy was done according to Priester et al29 with slight modifications. Cells were harvested by centrifugation (4°C, 10,000 x g, 6 min) and pellet was suspended in 0.1 M phosphate buffer saline (PBS, pH 7.2), suspended cells were centrifuged (4°C, 10,000 x g, 6 min), PBS was discarded and the cells were pre-fixed in mixture of glutaraldehyde (2.4% final concentration), ruthenium red (0.01%) and incubated for 30 min. After pre-fixation, cells were removed by centrifugation (4°C, 10,000 x g, 6 min) and suspended in 2.4% glutaraldehyde and 0.01% ruthenium red for overnight at 4°C. Later, cells were removed by centrifugation (4°C, 10,000 x g, 6 min) from fixation solution, washed in PBS thrice and finally post-fixed in 1% osmium tetroxide solution for 2 h at 4 °C. After postfixation, sample was washed with PBS and dehydrated by a series of ethanolic washes starting from 20%, 30%, 50%, 70%, 90% and 100% (v/v) ethanol. After dehydration cells were mounted on glass pieces (0.5x0.5 cm) and dried in a critical point dryer using standard protocol, dried samples were fixed to SEM stubs and coated with gold. The specimens were examined by using SEM (Philips XL30 series) at different magnification ranges. Isolation of R. benzoatilyticus JA2 proteome Experimental design contained three independent experiments, each with three biological replicates and to minimize the biological variations all three biological replicates of an individual experiment were pooled together. In case of control, we pooled proteome from three independent experiments into one. Aniline exposed and control cells were harvested by centrifugation (4°C, 10,000 x g, 10 min) and the cell pellet was suspended in 50 8

ACS Paragon Plus Environment

Page 9 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

mM HEPES-KOH buffer (pH 7.5) and washed twice with the same buffer. Finally, cells were re-suspended in 3 ml of the HEPES-KOH buffer containing 0.1% SDS (w/v) and 0.1% tritonX-100(v/v) and sonicated (MS 70 probe, 45% power, 7 cycles, 4°C) to lyse the cells. Lysate was incubated for 30 min at 4 °C, centrifuged (4°C, 20,000 x g for 30 min) and the clear supernatant was taken as soluble proteome. Total proteins were precipitated by six volumes of pre-chilled acetone at -20°C (1: 6) overnight and precipitated proteins were centrifuged (4°C, 10,000 x g for 15 min), washed with 100% acetone twice. Then acetone was decanted and protein was lyophilized and stored at -20°C till analysis. Isobaric tag relative and absolute quantitation (iTRAQ) labelling of proteome Isobaric tag relative and absolute quantitation analysis was outsourced from California University, USA and the following protocol was adopted. One hundred fifty micrograms of each sample was re-suspended in TNE buffer [50 mM Tris pH 8.0, 100 mM NaCl, 1 mM EDTA]. RapiGest SF reagent (Waters) was added to the mix to a final concentration of 0.1%. Samples were then heated for 5 min at 95°C. Proteins were reduced with 1 mM Tris-(2-carboxyethyl) phosphine (TCEP) (Pierce Chemical) for 30 min at 37°C and carboxymethylated with 0.5 mg/ml of iodoacetamide for 30 min at 37°C. Iodoacetamide was then neutralized with an additional 1 mM TCEP, proteins were digested with trypsin [trypsin ratio 1:100 (trypsin: protein)] overnight at 37°C.. Samples were then treated with 50 mM HCl at 37°C for 1 hour, followed by centrifugation at maximum speed for 30 min at 4°C to degrade and remove the RapiGest. The soluble fraction was then removed to a new tube and the pH of the solution was adjusted to 3.0 using NH4OH. The peptides were then extracted and desalted using Aspire RP30 Desalting Tips (Thermo Scientific). Peptides were re-quantified using bicinchonic acid assay and 100 µg of each sample was labelled with a unique iTRAQ tag (114, 115, 116, and 117) as described in the manufacturer’s protocol (ABSCIEX). Control sample was labelled with the number 114 and three aniline exposed 9

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 54

samples with 115,116 and 117, respectively. The 4 samples were then combined and dried down to remove ethanol using a speed-vac. The samples were re-suspended in 100 µl of Buffer A (98% H2O, 2% ACN, 0.2% formic acid, and 0.005% TFA) and 5 µl sample was used for the MudPIT (multidimensional protein identification technique) analysis. Two dimensional Nano LC-ESI-MS/MS analysis iTRAQ labelled peptide mixture was analysed on a

QSTAR-Elite hybrid mass

spectrometer (Applied Biosytems/MDS Sciex) interfaced to the nano-flow HPLC. Peptides were separated using nano-flow high pressure liquid chromatography (HPLC) coupled with tandem mass spectroscopy (LC-MS/MS) using nano-spray ionization source. Strong cation exchange (SCX) fractionation was carried on BioX-SCX (5 µm particle size, 0.5 mm inner diameter x 15 mm. LC Packings P/N 161395) trap column. The sample was loaded onto the SCX column and eluted with 7.5 µl/min flow rate for 10 min using the gradient of buffer A and buffer C (5% ACN, 0.2% formic acid, and 0.5 M ammonium acetate) The SCX salt steps (first dimension) used for separation were 5%, 7.5%, 10%, 12.5%, 15%, 20%, 25%, 30%, 40%, 50%, 75% and 100% (w/v). In the second dimension (reverse phase) peptides were eluted from the ZorbaxTM C18 column (100 x 0.18 mm, 5-µm, Agilent Technologies, Santa Clara, CA) into the mass spectrometer using a linear gradient of 5–80% Buffer B (100% ACN, 0.2% formic acid, and 0.005% TFA) and Buffer A (98% H2O, 2% ACN, 0.2% formic acid, and 0.005% TFA) over 60 min at 400 nl/min flow rate. LC-MS/MS data was acquired in a data-dependent fashion by selecting the 6 most intense peaks with charge state of plus 2 to 4 that exceeds 35 counts, with exclusion of former target ions set to "60 seconds" and the mass tolerance for exclusion set to 100 ppm. Time-of-flight MS were acquired at m/z 400 to 2000 Da for 0.75 sec with 12 time bins to sum. MS/MS data were acquired from m/z 50 to 2,000 Da by using "enhance all" and 24 time bins to sum, dynamic background subtract, automatic collision energy, and 10

ACS Paragon Plus Environment

Page 11 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

automatic MS/MS accumulation with the fragment intensity multiplier set to 6 and maximum accumulation set to 2 s before returning to the survey scan. Mass spectral data analysis, protein identification and quantification The data was acquired using Analyst QS 2.0 software (Applied Biosystems/MDS Sciex). The peak list generation, protein identification, and peptide quantification were performed using ProteinPilot v3.0 (Applied Biosystems MDS-Sciex, USA) with default parameters. Data was combined into a single search for identification and quantification. The database search was performed against the genome project R. benzoatilyticus JA2 database (version AEWG00000000.1,) including 3,947 predicted genes. The Paragon algorithm in the ProteinPilot software was used for the peptide identification and further processed by the Pro Group algorithm where isoform-specific quantification was adopted to trace the differences between expressions of various isoforms. The defined parameters were as follows: (i) Sample type, iTRAQ 4-plex (Peptide Labelled), lysine and N-terminus modified iTRAQ tag; (ii) Cysteine alkylation, Iodoacetamide; (iii) Digestion, Trypsin with a maximum of one missed cleavage allowed; (iv) Instrument, QSTAR Elite ESI; (v) Special factors, none; (vi) Species, none; (vii) Specify Processing, Quantitate, bias correction; (viii) ID Focus, Biological modifications, amino acid modifications; (ix) Search effort, thorough. The default precursors and fragment mass tolerances for QSTAR ESI MS instrument were adopted by the software. The peak areas and the S/N ratios were extracted from the database by Protein Pilot 3.0 in order to process the raw data to yield quantification data. The peptide for quantification was automatically selected by Pro Group algorithm criteria (identified with higher confidence 99%, the peptide was not shared with another protein, iTRAQ reporter area is not zero) to calculate the reporter peak area, error factor (EF) and p-value. The resulting data set was auto bias-corrected to get rid of any variations imparted because of the unequal mixing during combining different labelled samples. To minimize false positive results, a strict cut off for 11

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 54

protein identification was applied. Proteins that met the criteria ; unused protscore >2 (confidence level 99%) with two unique peptides (with 95% confidence) were considered in protein identification. The false discovery rate (FDR) of the identified protein (false discovery rate = [decoy hits/total hits] 100%) was estimated based on the decoy search strategy and plotted in the protein summary report (Supplementary data set). Approximately 756 proteins were identified with unused protscore > 2 with corresponding FDR 2, p ≤0.05, EF ≤2, FDR 1.35 were up-regulated.30, 31 iTRAQ reporter ratios were log-transformed (log10), mean and standard deviations were calculated from three independent experiments and consisted only of proteins with p ≤0.05 discovered in at least two out of three experiments. P-values of differential regulated proteins were subjected to multiple testing corrections by Benjamini and Hochberg method using an open-source software32 with significance level P ≤0.05.

Proteins were functionally

categorised according to KEGG (Kyoto Encyclopedia of Genes and Genomes, www.genome.jp/kegg/) database and in silico analysis of proteins was carried out by using ExPASy tools (www.expasy.org). Extraction and quantification of PHAs Aniline exposed (after 48 h of aniline exposure) or control cells were harvested by centrifugation (10,000 x g for 10 min, 4○C) and the cell pellet was washed twice with distilled water. Finally, cell pellet was re-suspended in distilled water and freeze dried (Labcanco USA). Freeze dried biomass (300 mg) was suspended in 30 ml of chloroform and incubated overnight in rotary shaker (120 rpm) at 30○C. Obtained suspension was filtered through Whatman No1 paper to remove cell debris and the filtrate was subjected to precipitation overnight at -20○C in 10 volumes of ethanol. Finally, obtained precipitate was concentrated by centrifugation (15,000 x g for 20 min) and dissolved in 5 ml of acetone. 12

ACS Paragon Plus Environment

Page 13 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

PHAs were purified by adding 20 ml of 70% ethanol (v/v), methanol in 1:1 ratio to PHAs suspended in acetone. Finally, pure PHAs were obtained by centrifugation (15,000 x g for 20 min), purified PHAs were air dried, sealed stored at -20○C. Indoles were estimated by Salper’s method using indole as standard33 and total sugars by the phenol-sulphuric acid method using glucose as standard34. Proteins were estimated by dye binding assay using BSA as standard.35 FAME analysis Fatty acids of the samples were identified by FAME analysis done at Royal Life Science Pvt Ltd. Fatty acids were extracted from lyophilized cells and fatty acid isolation, identification was done MIDI-MIS method36. First fatty acids were saponified, methylated, and extracted by using the protocol of the Sherlock microbial identification system (MIDI Inc.) procedure. The fatty acid methyl esters thus obtained were analyzed by gas chromatography equipped with Sherlock MIS software [Microbial ID; MIDI 6.0 version; Agilent: 6850; peak identification was done based on RTSBA6 database]. Results Growth and aniline tolerance Strain JA2 could not grow at the expense of aniline as sole carbon or nitrogen source. . However, strain JA2 was able to grow in the presence of high concentrations (5-25 mM) of aniline (Figure 1A) with minimum inhibition concentration (MIC) of 30 mM and IC50 of 23 mM. When mid log phase cultures were exposed to high concentrations aniline (25 mM), ~80% cells were viable (data not shown) which is in agreement with previous reports.24 To elucidate the responses of strain JA2 to aniline, 25 mM of aniline was added to the mid log phase cultures for all further experiments. SEM analysis of the aniline exposed cells revealed rough and altered cell surface with extracellular depositions as compared to the control (unexposed culture) (Figure 1B, C) indicating altered cell surface due to aniline stress. 13

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 54

Aniline transformation studies Although strain JA2 did not grow at the expense of aniline, loss of aniline (1-1.5%) in the presence of culture indicated a possible biotransformation of aniline. To identify the aniline transformation products, deuterium labelled aniline (aniline-d5, Sigma-Aldrich) feeding experiments were performed.

LC-MS analysis of

labelled aniline fed culture

supernatants revealed two aniline bio-transformed products, compound I (Rt11.58 min) and compound II (Rt15.58 min) which were absent in control (unexposed) cultures (Figure 2A). Compound I had a molecular ion mass of 137 [M+H] from unlabelled fraction and 142[M+H] in labelled fraction (Supplementary figure 1A,B). Compound II had a mass of 136 [M+H] from unlabelled fraction (Figure 2B) and 141[M+H] (Figure 2C) from labelled fraction. Five mass units (5 amu) increase in molecular ion masses of both the compounds (137 to 142; 136 to 141) suggest that these compounds were derived from aniline. Mass spectrum of compound II was (136 [M+H], 94 m/z) was identical to that of authentic acetanilide mass spectrum (mass bank; www.massbank.jp) and it was co-eluted with acetanilide standard, confirming that compound II is indeed acetanilide. Only ~0.1 mM of aniline was transformed to acetanilide and compound I could not be identified due to lack of spectral similarity in the database. Metabolic responses (adaptations) to aniline stress GC-MS analysis of intracellular metabolites of control and aniline exposed cells of R. benzoatilyticus JA2 revealed a total of 161 metabolic features of which 61 metabolic features were identified based on data base (NIST similarity >800, Glom data base), by comparing to the mass spectra of authentic standards (sugars, amino acids and organic acid, aromatic acids) and 100 metabolic features remained unidentified. Identified metabolites (61) include aliphatic organic acids, amino acids and their derivatives, sugars, nucleotides, fatty acids, amines and aromatic acids. Identified metabolites were used for multivariate statistical analysis to identify significant metabolic variations and metabolic patterns. Hierarchical 14

ACS Paragon Plus Environment

Page 15 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

clustering analysis (HCA) of metabolites from control and aniline exposed cells resulted in separation of the two groups. All aniline exposed samples clustered as a single distinct group; likewise all control samples also clustered as a single group (Supplementary figure 2A). HCA analysis indicated metabolic dissimilarity between control and aniline exposed cells. To identify significant changes in metabolism and metabolic signatures associated to aniline stress, data was analysed by robust statistical methods. Five component PCA analyses of the data set explained 90% of the variance, 74.1% of the variation was explained by principal component1 (PC1) and 10.4% by principle component 2 (PC2). Control and aniline exposed samples were clearly separated from each other (Figure 3A) indicating metabolic differences in control and aniline exposed cells. However, sub-clustering of control groups in the PCA plot was largely due to variations in relative levels of metabolites (leucine, serine, alanyl-glycine, phosphoglyceric acid, gluconic acid, trehalose, erythrose, thymidine and ethanolamine) within the control group. Though the experimental conditions were identical, variation in metabolite levels within control groups is plausible due to intrinsic biological variations within samples. Nevertheless, PCA analysis of control and aniline exposed samples (Figure. 3) suggested that aniline exposed cells are metabolically different from the control. Further, partial least square discrimination (PLS-DA) analysis also separated control and aniline exposed groups (Figure 3B) strongly suggesting metabolic variations between groups. R2 value for this model was 0.9 which indicated goodness of fit and Q2 was 0.85 which indicates the goodness of predictability. High values of R2 and Q2 indicate that this model is representative of true differences in metabolism. Key metabolic pathways modulated by aniline stress Variable importance on projection (VIP) scores obtained from PLS-DA model was used to identify key metabolic features significant for group separation and VIP scores. Metabolites with VIP score >1 were considered to have a statistically significant contribution 15

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 54

to the model. Thirty one metabolites were identified as statistically significant (VIP >1) from VIP score plot (Supplementary figure 2B) which were largely responsible for group separation in the model. Further, metabolites with VIP score >1 were selected and subjected to student’s t test to identify significantly regulated metabolites to aniline stress. Thirty one metabolites were identified whose response was altered significantly (fold change >2 and Pvalue ≤0.05) (Figure 4A). The relative concentration of 21 metabolites was high in aniline exposed cells while concentration of 10 metabolites were low compared to that of control (Figure 4A). Further, differentially regulated metabolites (identified in student’s t test) were annotated to their respective metabolic pathways according to the KEGG database (www.genome.jp/kegg) to identify metabolic pathways affected by aniline stress. Amino acid metabolism was highly (29%) affected by aniline stress followed by carbohydrate metabolism (14%), fatty acid metabolism (12%).

Butanoate (6%), nucleic acid (6%),

nitrogen (6%) and vitamins-cofactors metabolism and tricarboxylic acid cycle (6%) were also affected by aniline stress (Figure 4B). Proteomic inventory of Rubrivivax benzoatilyticus JA2 to aniline stress Isobaric Tag Relative and Absolute Quantification (iTRAQ) based comparative proteomic analysis was performed to decipher the proteomic responses of R. benzoatilyticus JA2 to aniline stress (Supplementary figure 3A). Seven hundred and fifty six proteins were identified (Supplementary table 1) after applying the cut off of the unused protein score > 2, EF ≤ 2 and a false discovery rate of < 1.0% (Supplementary data set.), which correspond to 16% of the total theoretical proteome (3,898 protein coding genes of R. benzoatilyticus JA2). All the 756 proteins were detected in all three independent experiments (Supplementary table 1) and linear regression analysis (fold changes) of the iTRAQ identified proteins of two biological replicates which had a correlation coefficient value (R, R2) of R=0.84, R2= 0.721 16

ACS Paragon Plus Environment

Page 17 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

which indicate that the data was highly reproducible (Supplementary figure 3B). Identified proteins were subjected to in silico analysis for theoretical molecular weight, isoelectric point (pI) and hydropathy analysis (www.expasy.org). Molecular weight versus isoelectric point map of identified proteins indicated two clusters of proteins with pI of 4.5- 7.0 and 9.0-11, respectively (Supplementary figure 4A). Of the total proteins identified, grand average hydropathy (GRAVY) analysis indicated 31.5% to be hydrophilic and 68.5% as hydrophobic proteins (Supplementary figure 4B). Proteins identified (756) by iTRAQ analysis were subjected to volcano plot analysis to identify differentially regulated proteins. A protein was considered differentially regulated if the fold change >1.35 (aniline exposed/control) in at least two biological replicates with p ≤0.05. A total of 114 proteins were identified as differentially regulated (Figure 5A) of which 58 proteins were up-regulated (aniline exposed/control ratio >1.35) and 56 proteins were down regulated (ratio 2, fold change of >1.35, Pvalue ≤0.05 (raw and multiple testing corrected) were presented in table and proteins with Pvalue > 0.05 are in italics. Proteins shaded in grey indicate up-regulated and un-shaded are down regulated.

41

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

References 1. Skipper, P. L.; Kim, M. Y.; Sun, H. L.; Wogan, G. N.; Tannenbaum, S. R., Monocyclic aromatic amines as potential human carcinogens: old is new again. Carcinogenesis 2010, 31, (1), 50-8. 2. Martins, M.; Rodrigues-Lima, F.; Dairou, J.; Lamouri, A.; Malagnac, F.; Silar, P.; Dupret, J. M., An acetyltransferase conferring tolerance to toxic aromatic amine chemicals: molecular and functional studies. J Biol Chem 2009, 284, (28), 18726-33. 3. Cocaign, A.; Bui, L. C.; Silar, P.; Chan Ho Tong, L.; Busi, F.; Lamouri, A.; Mougin, C.; RodriguesLima, F.; Dupret, J. M.; Dairou, J., Biotransformation of Trichoderma spp. and their tolerance to aromatic amines, a major class of pollutants. Appl Environ Microbiol 2013, 79, (15), 4719-26. 4. Kim, D.; Guengerich, F. P., Cytochrome P450 activation of arylamines and heterocyclic amines. Annu Rev Pharmacol Toxicol 2005, 45, 27-49. 5. Liang, Q.; Takeo, M.; Chen, M.; Zhang, W.; Xu, Y.; Lin, M., Chromosome-encoded gene cluster for the metabolic pathway that converts aniline to TCA-cycle intermediates in Delftia tsuruhatensis AD9. Microbiology 2005, 151, (Pt 10), 3435-46. 6. Vangnai, A. S.; Petchkroh, W., Biodegradation of 4-chloroaniline by bacteria enriched from soil. FEMS Microbiol Lett 2007, 268, (2), 209-16. 7. Kahng, H. Y.; Kukor, J. J.; Oh, K. H., Characterization of strain HY99, a novel microorganism capable of aerobic and anaerobic degradation of aniline. FEMS Microbiol Lett 2000, 190, (2), 215-21. 8. Liu, Z.; Yang, H.; Huang, Z.; Zhou, P.; Liu, S. J., Degradation of aniline by newly isolated, extremely aniline-tolerant Delftia sp. AN3. Appl Microbiol Biotechnol 2002, 58, (5), 679-82. 9. Takenaka, S.; Mulyono; Sasano, Y.; Takahashi, Y.; Murakami, S.; Aoki, K., Microbial transformation of aniline derivatives: regioselective biotransformation and detoxification of 2phenylenediamine by Bacillus cereus strain PDa-1. J Biosci Bioeng 2006, 102, (1), 21-7. 10. Schnell, S.; Bak, F.; Pfennig, N., Anaerobic degradation of aniline and dihydroxybenzenes by newly isolated sulfate-reducing bacteria and description of Desulfobacterium anilini. Arch Microbiol 1989, 152, (6), 556-63. 11. Tao, F.; Tang, H.; Gai, Z.; Su, F.; Wang, X.; He, X.; Xu, P., Genome sequence of Pseudomonas putida Idaho, a unique organic-solvent-tolerant bacterium. J Bacteriol 2011, 193, (24), 7011-2. 12. Sardessai, Y. N.; Bhosle, S., Industrial potential of organic solvent tolerant bacteria. Biotechnol Prog 2004, 20, (3), 655-60. 13. Rodriguez-Herva, J. J.; Garcia, V.; Hurtado, A.; Segura, A.; Ramos, J. L., The ttgGHI solvent efflux pump operon of Pseudomonas putida DOT-T1E is located on a large self-transmissible plasmid. Environ Microbiol 2007, 9, (6), 1550-61. 14. Udaondo, Z.; Duque, E.; Fernandez, M.; Molina, L.; Torre Jde, L.; Bernal, P.; Niqui, J. L.; Pini, C.; Roca, A.; Matilla, M. A.; Antonia Molina-Henares, M.; Silva-Jimenez, H.; Navarro-Aviles, G.; Busch, A.; Lacal, J.; Krell, T.; Segura, A.; Ramos, J. L., Analysis of solvent tolerance in Pseudomonas putida DOT-T1E based on its genome sequence and a collection of mutants. FEBS Lett 2012, 586, (18), 29328. 15. Puchalka, J.; Oberhardt, M. A.; Godinho, M.; Bielecka, A.; Regenhardt, D.; Timmis, K. N.; Papin, J. A.; Martins dos Santos, V. A., Genome-scale reconstruction and analysis of the Pseudomonas putida KT2440 metabolic network facilitates applications in biotechnology. PLoS Comput Biol 2008, 4, (10), e1000210. 16. Dominguez-Cuevas, P.; Gonzalez-Pastor, J. E.; Marques, S.; Ramos, J. L.; de Lorenzo, V., Transcriptional tradeoff between metabolic and stress-response programs in Pseudomonas putida KT2440 cells exposed to toluene. J Biol Chem 2006, 281, (17), 11981-91. 17. Naether, D. J.; Slawtschew, S.; Stasik, S.; Engel, M.; Olzog, M.; Wick, L. Y.; Timmis, K. N.; Heipieper, H. J., Adaptation of the hydrocarbonoclastic bacterium Alcanivorax borkumensis SK2 to alkanes and toxic organic compounds: a physiological and transcriptomic approach. Appl Environ Microbiol 2013, 79, (14), 4282-93. 18. Trautwein, K.; Kuhner, S.; Wohlbrand, L.; Halder, T.; Kuchta, K.; Steinbuchel, A.; Rabus, R., Solvent stress response of the denitrifying bacterium "Aromatoleum aromaticum" strain EbN1. Appl Environ Microbiol 2008, 74, (8), 2267-74. 42

ACS Paragon Plus Environment

Page 42 of 54

Page 43 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

19. Segura, A.; Godoy, P.; van Dillewijn, P.; Hurtado, A.; Arroyo, N.; Santacruz, S.; Ramos, J. L., Proteomic analysis reveals the participation of energy- and stress-related proteins in the response of Pseudomonas putida DOT-T1E to toluene. J Bacteriol 2005, 187, (17), 5937-45. 20. Yun, S. H.; Park, G. W.; Kim, J. Y.; Kwon, S. O.; Choi, C. W.; Leem, S. H.; Kwon, K. H.; Yoo, J. S.; Lee, C.; Kim, S.; Kim, S. I., Proteomic characterization of the Pseudomonas putida KT2440 global response to a monocyclic aromatic compound by iTRAQ analysis and 1DE-MudPIT. J Proteomics 2011, 74, (5), 620-8. 21. Puglisi, E.; Cahill, M. J.; Lessard, P. A.; Capri, E.; Sinskey, A. J.; Archer, J. A.; Boccazzi, P., Transcriptional response of Rhodococcus aetherivorans I24 to polychlorinated biphenylcontaminated sediments. Microb Ecol 2010, 60, (3), 505-15. 22. Ni, Y.; Song, L.; Qian, X.; Sun, Z., Proteomic analysis of Pseudomonas putida reveals an organic solvent tolerance-related gene mmsB. PLoS One 2013, 8, (2), e55858. 23. Ramana Ch, V.; Sasikala, C.; Arunasri, K.; Anil Kumar, P.; Srinivas, T. N.; Shivaji, S.; Gupta, P.; Suling, J.; Imhoff, J. F., Rubrivivax benzoatilyticus sp. nov., an aromatic, hydrocarbon-degrading purple betaproteobacterium. Int J Syst Evol Microbiol 2006, 56, (Pt 9), 2157-64. 24. Mujahid, M.; Sasikala, C.; Ramana Ch, V., Aniline-induced tryptophan production and identification of indole derivatives from three purple bacteria. Curr Microbiol 2010, 61, (4), 285-90. 25. Kalyan Chakravarthy, S.; Srinivas, T. N.; Anil Kumar, P.; Sasikala, C.; Ramana Ch, V., Roseospira visakhapatnamensis sp. nov. and Roseospira goensis sp. nov. Int J Syst Evol Microbiol 2007, 57, (Pt 11), 2453-7. 26. Gromova, M.; Roby, C., Toward Arabidopsis thaliana hydrophilic metabolome: assessment of extraction methods and quantitative 1H NMR. Physiol Plant 2010, 140, (2), 111-27. 27. Jozefczuk, S.; Klie, S.; Catchpole, G.; Szymanski, J.; Cuadros-Inostroza, A.; Steinhauser, D.; Selbig, J.; Willmitzer, L., Metabolomic and transcriptomic stress response of Escherichia coli. Mol Syst Biol 2010, 6, 364. 28. Xia, J.; Psychogios, N.; Young, N.; Wishart, D. S., MetaboAnalyst: a web server for metabolomic data analysis and interpretation. Nucleic Acids Res 2009, 37, (Web Server issue), W65260. 29. Priester, J. H.; Horst, A. M.; Van de Werfhorst, L. C.; Saleta, J. L.; Mertes, L. A.; Holden, P. A., Enhanced visualization of microbial biofilms by staining and environmental scanning electron microscopy. J Microbiol Methods 2007, 68, (3), 577-87. 30. Ruppen, I.; Grau, L.; Orenes-Pinero, E.; Ashman, K.; Gil, M.; Algaba, F.; Bellmunt, J.; SanchezCarbayo, M., Differential protein expression profiling by iTRAQ-two-dimensional LC-MS/MS of human bladder cancer EJ138 cells transfected with the metastasis suppressor KiSS-1 gene. Mol Cell Proteomics 2010, 9, (10), 2276-91. 31. Chen, Y.; Choong, L. Y.; Lin, Q.; Philp, R.; Wong, C. H.; Ang, B. K.; Tan, Y. L.; Loh, M. C.; Hew, C. L.; Shah, N.; Druker, B. J.; Chong, P. K.; Lim, Y. P., Differential expression of novel tyrosine kinase substrates during breast cancer development. Mol Cell Proteomics 2007, 6, (12), 2072-87. 32. Lesack, K.; Naugler, C., An open-source software program for performing Bonferroni and related corrections for multiple comparisons. J Pathol Inform 2011, 2, 52. 33. Powell, L. E., Preparation of Indole Extracts from Plants for Gas Chromatography and Spectrophotofluorometry. Plant Physiol 1964, 39, (5), 836-42. 34. Jiao, Y.; Cody, G. D.; Harding, A. K.; Wilmes, P.; Schrenk, M.; Wheeler, K. E.; Banfield, J. F.; Thelen, M. P., Characterization of extracellular polymeric substances from acidophilic microbial biofilms. Appl Environ Microbiol 2010, 76, (9), 2916-22. 35. Bradford, M. M., A rapid and sensitive method for the quantitation of microgram quantities of protein utilizing the principle of protein-dye binding. Anal Biochem 1976, 72, 248-54. 36. Sasser, M., Identification of bacteria by gas chromatography of cellular fatty acids. MIDI Inc: Newark 1997; Vol. MIDI Technical Note 101. 37. Ramos, J. L.; Duque, E.; Gallegos, M. T.; Godoy, P.; Ramos-Gonzalez, M. I.; Rojas, A.; Teran, W.; Segura, A., Mechanisms of solvent tolerance in gram-negative bacteria. Annu Rev Microbiol 2002, 56, 743-68. 43

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 44 of 54

38. Ye, Y.; Zhang, L.; Hao, F.; Zhang, J.; Wang, Y.; Tang, H., Global metabolomic responses of Escherichia coli to heat stress. J Proteome Res 2012, 11, (4), 2559-66. 39. Auger, C.; Lemire, J.; Cecchini, D.; Bignucolo, A.; Appanna, V. D., The metabolic reprogramming evoked by nitrosative stress triggers the anaerobic utilization of citrate in Pseudomonas fluorescens. PLoS One 2011, 6, (12), e28469. 40. Ye, Y.; Wang, X.; Zhang, L.; Lu, Z.; Yan, X., Unraveling the concentration-dependent metabolic response of Pseudomonas sp. HF-1 to nicotine stress by (1)H NMR-based metabolomics. Ecotoxicology 2012, 21, (5), 1314-24. 41. Miller, C. D.; Pettee, B.; Zhang, C.; Pabst, M.; McLean, J. E.; Anderson, A. J., Copper and cadmium: responses in Pseudomonas putida KT2440. Lett Appl Microbiol 2009, 49, (6), 775-83. 42. Mujahid, M.; Sasikala, C.; Ramana Ch, V., Aniline is an inducer, and not a precursor, for indole derivatives in Rubrivivax benzoatilyticus JA2. PLoS One 2014, 9, (2), e87503. 43. Shen, T.; Liu, Q.; Xie, X.; Xu, Q.; Chen, N., Improved production of tryptophan in genetically engineered Escherichia coli with TktA and PpsA overexpression. J Biomed Biotechnol 2012, 2012, 605219. 44. Gosset, G., Production of aromatic compounds in bacteria. Curr Opin Biotechnol 2009, 20, (6), 651-8. 45. Hirasawa, T.; Yoshikawa, K.; Nakakura, Y.; Nagahisa, K.; Furusawa, C.; Katakura, Y.; Shimizu, H.; Shioya, S., Identification of target genes conferring ethanol stress tolerance to Saccharomyces cerevisiae based on DNA microarray data analysis. J Biotechnol 2007, 131, (1), 34-44. 46. Han, T. H.; Lee, J. H.; Cho, M. H.; Wood, T. K.; Lee, J., Environmental factors affecting indole production in Escherichia coli. Res Microbiol 2011, 162, (2), 108-16. 47. Mittal, M.; Rockne, K. J., Indole production by Pseudomonas stutzeri strain NAP-3 during anaerobic naphthalene biodegradation in the presence of dimethyl formamide. J Environ Sci Health A Tox Hazard Subst Environ Eng 2008, 43, (9), 1027-34. 48. Lee, J. H.; Lee, J., Indole as an intercellular signal in microbial communities. FEMS Microbiol Rev 2010, 34, (4), 426-44. 49. Mrozik, A.; Piotrowska-Seget, Z.; Labuzek, S., Changes in whole cell-derived fatty acids induced by naphthalene in bacteria from genus Pseudomonas. Microbiol Res 2004, 159, (1), 87-95. 50. York, G. M.; Junker, B. H.; Stubbe, J. A.; Sinskey, A. J., Accumulation of the PhaP phasin of Ralstonia eutropha is dependent on production of polyhydroxybutyrate in cells. J Bacteriol 2001, 183, (14), 4217-26. 51. Susana Castro-Sowinski; Saul Burdman; Ofra Matan; Okon, Y., Plastics from bacteria: Natural functions of bacterial polyhydroxyalkanoates. Springer-Verlag: Heidelberg, Germany, 2010; Vol. 14, p 39-61. 52. Zhao, Y. H.; Li, H. M.; Qin, L. F.; Wang, H. H.; Chen, G. Q., Disruption of the polyhydroxyalkanoate synthase gene in Aeromonas hydrophila reduces its survival ability under stress conditions. FEMS Microbiol Lett 2007, 276, (1), 34-41. 53. Halan, B.; Schmid, A.; Buehler, K., Real-time solvent tolerance analysis of Pseudomonas sp. strain VLB120{Delta}C catalytic biofilms. Appl Environ Microbiol 2011, 77, (5), 1563-71. 54. White, A. P.; Weljie, A. M.; Apel, D.; Zhang, P.; Shaykhutdinov, R.; Vogel, H. J.; Surette, M. G., A global metabolic shift is linked to Salmonella multicellular development. PLoS One 2010, 5, (7), e11814. 55. Cabral, M. P.; Soares, N. C.; Aranda, J.; Parreira, J. R.; Rumbo, C.; Poza, M.; Valle, J.; Calamia, V.; Lasa, I.; Bou, G., Proteomic and functional analyses reveal a unique lifestyle for Acinetobacter baumannii biofilms and a key role for histidine metabolism. J Proteome Res 2011, 10, (8), 3399-417. 56. Barreto, M.; Jedlicki, E.; Holmes, D. S., Identification of a gene cluster for the formation of extracellular polysaccharide precursors in the chemolithoautotroph Acidithiobacillus ferrooxidans. Appl Environ Microbiol 2005, 71, (6), 2902-9. 57. Schlictman, D.; Shankar, S.; Chakrabarty, A. M., The Escherichia coli genes sspA and rnk can functionally replace the Pseudomonas aeruginosa alginate regulatory gene algR2. Mol Microbiol 1995, 16, (2), 309-20. 44

ACS Paragon Plus Environment

Page 45 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

58. Shankar, S.; Schlictman, D.; Chakrabarty, A. M., Regulation of nucleoside diphosphate kinase and an alternative kinase in Escherichia coli: role of the sspA and rnk genes in nucleoside triphosphate formation. Mol Microbiol 1995, 17, (5), 935-43. 59. Fang, H. H.; Xu, L. C.; Chan, K. Y., Effects of toxic metals and chemicals on biofilm and biocorrosion. Water Res 2002, 36, (19), 4709-16. 60. Flemming, H. C.; Neu, T. R.; Wozniak, D. J., The EPS matrix: the "house of biofilm cells". J Bacteriol 2007, 189, (22), 7945-7. 61. Sadykov, M. R.; Zhang, B.; Halouska, S.; Nelson, J. L.; Kreimer, L. W.; Zhu, Y.; Powers, R.; Somerville, G. A., Using NMR metabolomics to investigate tricarboxylic acid cycle-dependent signal transduction in Staphylococcus epidermidis. J Biol Chem 2010, 285, (47), 36616-24. 62. Mascher, T.; Helmann, J. D.; Unden, G., Stimulus perception in bacterial signal-transducing histidine kinases. Microbiol Mol Biol Rev 2006, 70, (4), 910-38. 63. Sanchez-Torres, V.; Maeda, T.; Wood, T. K., Global regulator H-NS and lipoprotein NlpI influence production of extracellular DNA in Escherichia coli. Biochem Biophys Res Commun 2010, 401, (2), 197-202. 64. Erol, I.; Jeong, K. C.; Baumler, D. J.; Vykhodets, B.; Choi, S. H.; Kaspar, C. W., H-NS controls metabolism and stress tolerance in Escherichia coli O157:H7 that influence mouse passage. BMC Microbiol 2006, 6, 72. 65. Hoffmann, A.; Bukau, B.; Kramer, G., Structure and function of the molecular chaperone Trigger Factor. Biochim Biophys Acta 2010, 1803, (6), 650-61. 66. Segura, A.; Molina, L.; Fillet, S.; Krell, T.; Bernal, P.; Munoz-Rojas, J.; Ramos, J. L., Solvent tolerance in Gram-negative bacteria. Curr Opin Biotechnol 2012, 23, (3), 415-21. 67. Moliere, N.; Turgay, K., Chaperone-protease systems in regulation and protein quality control in Bacillus subtilis. Res Microbiol 2009, 160, (9), 637-44. 68. Kobayashi, Y.; Ohtsu, I.; Fujimura, M.; Fukumori, F., A mutation in dnaK causes stabilization of the heat shock sigma factor sigma32, accumulation of heat shock proteins and increase in toluene-resistance in Pseudomonas putida. Environ Microbiol 2010, 13, (8), 2007-17. 69. Kruger, E.; Witt, E.; Ohlmeier, S.; Hanschke, R.; Hecker, M., The clp proteases of Bacillus subtilis are directly involved in degradation of misfolded proteins. J Bacteriol 2000, 182, (11), 325965. 70. Gerth, U.; Kruger, E.; Derre, I.; Msadek, T.; Hecker, M., Stress induction of the Bacillus subtilis clpP gene encoding a homologue of the proteolytic component of the Clp protease and the involvement of ClpP and ClpX in stress tolerance. Mol Microbiol 1998, 28, (4), 787-802. 71. Ozcan, N.; Ejsing, C. S.; Shevchenko, A.; Lipski, A.; Morbach, S.; Kramer, R., Osmolality, temperature, and membrane lipid composition modulate the activity of betaine transporter BetP in Corynebacterium glutamicum. J Bacteriol 2007, 189, (20), 7485-96. 72. Fernandes, P.; Ferreira, B. S.; Cabral, J. M., Solvent tolerance in bacteria: role of efflux pumps and cross-resistance with antibiotics. Int J Antimicrob Agents 2003, 22, (3), 211-6.

45

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 46 of 54

Figure for graphic abstract

46

ACS Paragon Plus Environment

Page 47 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

Figure 1

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2

ACS Paragon Plus Environment

Page 48 of 54

Page 49 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

Figure 3 242x138mm (300 x 300 DPI)

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 4

ACS Paragon Plus Environment

Page 50 of 54

Page 51 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

Figure 5

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 6

ACS Paragon Plus Environment

Page 52 of 54

Page 53 of 54

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of Proteome Research

Figure 7

ACS Paragon Plus Environment

Journal of Proteome Research

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 8 158x114mm (300 x 300 DPI)

ACS Paragon Plus Environment

Page 54 of 54