Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

System-wide analysis of manganese starvation-induced metabolism in key elements of Lactobacillus plantarum

Yanjun Tonga, Qixiao Zhaiad, Gang Wangad, Qiuxiang Zhanga, Xiaoming Liuad, Fengwei Tianad, Jianxin Zhaoacd, Hao Zhangacd and Wei Chen*abcd
aState Key Laboratory of Food Science and Technology, School of Food Science and Technology, Jiangnan University, Wuxi 214122, People's Republic of China. E-mail: tongyanjun@jiangnan.edu.cn; chenwei66@jiangnan.edu.cn
bBeijing Innovation Centre of Food Nutrition and Human Health, Beijing Technology & Business University, Beijing 100048, People's Republic of China
cInternational Joint Research Center for Probiotics & Gut Health, Jiangnan University, Wuxi 214122, P.R. China
dInternational Joint Research Laboratory for Probiotics at Jiangnan University, People's Republic of China

Received 3rd January 2017 , Accepted 10th February 2017

First published on 24th February 2017


Abstract

To analyze the response mechanisms of Lactobacillus plantarum against manganese starvation stress, different metabolisms from physiology, proteomics and transporters aspects in L. plantarum CCFM 436 were systematically investigated. The kinetics of cell growth (μmax) decreased from 0.310 to 0.256 h−1, while thinner cell morphology was observed by transmission electron microscopy under Mn-starvation conditions. Gas chromatography-mass spectrometry analysis indicated that membrane mobility and compactness increased, with a higher proportion of unsaturated fatty acids and cyclopropane fatty acids. High-performance liquid chromatography analysis showed that intracellular Asp, Glu, and Arg contents, closely related to energy metabolism, were significantly increased. Fourier transform infrared spectroscopy proved that some functional groups (N–H and O[double bond, length as m-dash]C–OH) were significantly affected by Mn starvation. Comparative two-dimensional proteomic analysis identified 73 proteins that differed significantly under Mn starvation conditions. These differentially expressed proteins involved in carbohydrate, amino acid and transcription/translation metabolisms and stress response were categorized as crucial components required to resist manganese starvation stress. Moreover, qRT-PCR analysis proved that MntH 1–5, negatively regulated by MntR, acted as potential Mn importers under Mn starvation conditions. The proposed coordinated mechanism model provides a reference for, and insight into, the intracellular metabolism of LAB strains.


1. Introduction

Manganese is an essential element of life and acts as “cell security” or a “life bodyguard”.1–3 Manganese is a component in a variety of enzymes with important physiological functions, such as manganese-superoxide dismutase, arginase and pyruvate carboxylase.4,5 Manganese also acts as a cofactor in some metalloenzymes, such as aminopeptidase, dipeptidase, manganese catalase, D-xylose isomerase, L-arabinose isomerase and ribozymes.3,4 Manganese must be maintained at an appropriate level in organisms; otherwise, stress can result from manganese starvation or excess.1 It is important to maintain a steady-state balance in the regulation of manganese to ensure normal physiological function.

Lactobacillus plantarum (L. plantarum) is a probiotic that can effectively reduce heavy metal ion toxicity.6–10 In particular, L. plantarum requires a relatively high level of manganese ions for optimal growth due to the absence of superoxide dismutase.11,12 Manganese is an essential element in the growth of lactic acid bacteria, not only involved in the composition of some enzymes, such as lactate dehydrogenase, but also promoting the production of certain enzymes to assist cellular metabolism.11 Manganese ions are indispensable to the stability and function of the metal protein, affecting the biological function of these proteins. These metalloproteins and metalloenzymes are involved in the regulation of homeostasis, electron transport, biomolecule synthesis, biological substance transport, oxidative stress, signal transduction, gene regulation and other functions.11,13,14 Therefore, the steady-state regulation of manganese is important in maintaining the normal physiological function of intracellular manganese ions. However, decreasing heavy metal ion toxicity has been the focus of most research in this area.15–18 Although manganese ion studies have focused on individual features, resistance to manganese starvation stress has not been studied systematically. Understanding the molecular mechanisms of the response to manganese starvation will be helpful in elucidating the physiological mechanisms of this kind of probiotic against manganese starvation stress from a different angle.

Our previous has shown the strong potential of L. plantarum CCFM436 to reduce manganese toxicity.19 Herein, L. plantarum CCFM436 was used to study metabolic response mechanisms under manganese starvation stress by investigating cell physiology, proteomics and levels of key transport proteins.

2. Materials and methods

2.1 Microbial growth conditions

L. plantarum CCFM436, which was screened from wine cake samples of Shanlan wine from Hainan, China, was cultured in modified MRS broth at 37 °C for 18 h (Table 1). The manganese ion concentration in the MRS broth was 16 mg L−1, while MRS broth without added MnSO4·H2O was used to induce manganese starvation stress. The biomass at different culture phases was determined as the optical density at 600 nm (OD600).
Table 1 Main characteristics of the targeted strain used in the study
Species Strain Microbial type Optimal T/pH Description and source
Lactobacillus plantarum CCFM436 Lactic acid bacteria; chemoheterotroph T 37 °C; pH 6.2–6.4 Wine cake sample of shanlan wine from Hainan, China


2.2 Analysis of cell physiology under manganese starvation conditions

2.2.1 Analysis of cell morphology using transmission electron microscopy (TEM). A 5 mL cultured medium was sampled and treated with glutaraldehyde at 4 °C for 0.5 h. The cells were centrifuged at 6000 × g for 5 min and then immobilized with agar. The agar-cell slice samples were soaked in phosphate buffer containing 2.5% glutaraldehyde and then transferred into osmium tetroxide solution. The slices were treated with uranyl acetate solution after being washed with double-distilled H2O (ddH2O). The slices were dehydrated with ethanol and propylene oxide, respectively, and then stained with uranyl acetate. After implanting the slices into epoxy, the cell morphology of the sample was observed with a transmission electron microscope (Hitachi-H 7000, Tokyo, Japan).
2.2.2 Analysis of membrane fatty acids composition by gas chromatography-mass spectrometry (GC-MS). Cells (1.0 g) were collected by centrifugation at 6000 × g for 5 min and dissolved in 1.0 M NaOH–methanol (2.5 mL) and incubated at 70 °C for 2 h. After cooling to room temperature, a solution of 25% BF3–methanol (2.5 mL) was added to the sample after cooling to room temperature, and the mixed sample was incubated at 65 °C for 2 h. The sample was again cooled at 20 °C for 10 min, and then 2 mL each of n-hexane and saturated NaCl solution was added sequentially and vortexed for 1 min. After centrifugation at 380 × g for 1 min, redundant water was removed with a moderate amount of anhydrous Na2SO4, and the supernatant oil was collected. The fatty acid composition was assayed using a trace GC-MS instrument (1200 L, Varian, Palo Alto, CA).

2.3 Analysis of intracellular amino acid composition and functional groups under manganese starvation conditions

2.3.1 Analysis of amino acid composition by high performance liquid chromatography (HPLC). A 50 mL sample of cultured medium was centrifuged at 8000 × g for 10 min. The collected thallus was washed three times with ddH2O and its weight recorded. The collected thallus was then treated with 5% trichloroacetic acid in a 25 mL volumetric flask at 25 °C for 1 h. The lysis solutions were filtrated and the mother liquor centrifuged at 12[thin space (1/6-em)]000 × g for 10 min, after which the supernatants were used to analyze the intracellular amino acid by HPLC.
2.3.2 Analysis of functional groups by Fourier transform infrared spectroscopy (FTIR). Different dried cell samples (approx. 2 mg) and KBr (200 mg) were repeatedly ground in an agate mortar. The sample was collected and turned into a thin cylinder using a press. The cylinder was measured in the wavelength region 400–4000 cm−1 using a Fourier transform infrared spectrometer (NEXUS, Nicolet, USA).

2.4 Two-dimensional (2D) gel electrophoresis

2.4.1 Isolation of whole proteins. Bacterial cells with different manganese concentrations were centrifuged at 6000 × g for 10 min and washed three times with phosphate buffer for proteomic study. Thalli were resuspended in enzymatic lysis buffer with mutanolysin (7 M urea, 2 M thiourea and 40% CHAPS, 40 mM Tris-base with 40 mM dithiothreitol, 1 mM phenylmethane sulfonyl fluoride, 2% immobilized pH gradient buffer (pH 3–10), protease inhibitor and nuclease mix) at 37 °C for 2 h and then sonicated for 15 min on ice. The lysate was then centrifuged at 12[thin space (1/6-em)]000 × g for 30 min at 4 °C so that the supernatants contained the whole protein from the samples. The Bradford assay was used to determine protein concentrations.
2.4.2 2D gel electrophoresis and image analysis. Samples containing proteins (1000 μg) were diluted in IPG strip rehydration buffer (9 M urea, 4% CHAPS, 1% immobilized pH gradient buffer (pH 3–10), 1% dithiothreitol and 0.002% bromophenol blue) and then loaded on IPG strips at pH 3–10. Isoelectric focusing was carried out at 20 °C on an IPGhor Isoelectric Focusing System using the optimal program. After focusing, the IPG strips were equilibrated at room temperature for 15 min in equilibration buffer (50 mmol L−1 Tris–HCl pH 8.8, 6 mol L−1 urea, 30% glycerol, 2% sodium dodecyl sulfate and 0.002% bromophenol blue) with the addition of 1% dithiothreitol, and for a further 15 min in the same equilibration buffer with the addition of 2.5% indole-3-acetic acid. The equilibrated strips were transferred into homogeneous 12.5% sodium dodecyl sulfate-polyacrylamide gels for 2D electrophoresis. After separation, the 2D electrophoresis gels were stained using Coomassie Blue G250.

The stained gels (three independent analytical replicates for each manganese concentration) were scanned at 300 dpi with an image scanner and analyzed with PDQuest software. Briefly, spot detection was carried out using optimized setting values for spot intensity, spot area and saliency, as determined by applying real-time filters to minimize artifact detection. After spot detection, manual spot editing was carried out to remove artifacts that escaped the filtering process. Spots that showed significant intensity differences (>1.5-fold or <0.67-fold change) between samples with a P value of <0.05 in different manganese concentrations were considered differentially expressed proteins.

2.4.3 MS analysis. Different protein spots were manually excised from the gels, washed with ddH2O and destained in buffer (100 mM NH4HCO3 in 30% acetonitrile) three times. Protein spots were then vacuum dried. All protein spots were digested with 50 ng trypsin in 30 μL buffer (25 mM NH4HCO3 with the addition of 10% acetonitrile) at 37 °C for 20 h. The supernatants were transferred into a new tube and then vacuum dried. The dried peptides were dissolved in TCA buffer (0.1% TCA, 0.7 mg mL−1 α-cyano-4-hydroxy-trans-cinnamic acid in acetonitrile/TCA; 85[thin space (1/6-em)]:[thin space (1/6-em)]0.1 v/v). All respective peptides were spotted on the sample target plate and analyzed by MALDI-TOF/TOF mass spectrometry. Peptides digested with trypsin were analyzed in positive ion mode, and the spectra were adjusted by peptide calibration. The spectra were analyzed using FlexAnalysis and BioTools software. The peptide results were compared and analyzed with the following parameters: NCBI-nr bacteria database, trypsin as the digestion enzyme, acceptance of cysteine carbamidomethylation, methionine oxidation, fragment mass tolerance of 0.99 Da and peptide mass tolerance of 300 ppm.

2.5 Quantitative RT-PCR

Quantitative RT-PCR analysis was used to quantify protein transcription. Total RNA was extracted from different samples with Trizol reagent according to manufacturer instructions. The RNA concentration was determined, and reverse transcription reactions were carried out with a PrimeScript RT Reagent Kit with gDNA Eraser. Then, 10 μL of a genome DNA removal mixture, containing 1 μg extracted total RNA, 2 μL 5 × DNA eraser buffer, 1 μL gDNA eraser and RNase-free dH2O, was added. The reaction was carried out at 42 °C for 2 min, and the RT-PCR reaction was processed using a 20 μL mixture (1 μL PrimeScript RT enzyme mix, 1 μL RT primer mix, 4 μL 5 × PrimeScript buffer, 4 μL RNase free dH2O, and 10 μL of the reaction mixture for genome DNA removal). qRT-PCR was performed using SYBR Premix Ex Taq II. The 20 μL PCR reaction mixture contained 10 μL SYBR Premix Ex Taq II, 1 μL PCR forward primer (10 μM), 1 μL PCR reverse primer (10 mM), 1 μL sample cDNA and 7 μL dH2O. PCR specificity was verified by melt-curve analysis of the samples. The cycle threshold (Ct) value was measured, and the relative quantification of specific gene expressions was determined using the 2−ΔΔCt method. Primers are listed in Table 2. 16S rRNA of L. plantarum CCFM436 was used as an internal control for PCR amplification.
Table 2 Primer sequences used for qRT-PCR experiments
Primer Sequence (5′-3′) Production (bp)
MntH-1-F TCACCGTGGCATACAGTGACACAC 187
MntH-1-R TGAAATTGTTTTAGCGCACGACCT
MntH-2-F TACGTGGCATGTCGATGATGG 203
MntH-2-R CACGACCGTCGTCGTCATCATATC
MntH-3-F TACATGGATCCCGGTAACTGGTCAAC 145
MntH-3-R GATCCATCTGACTCACGATACCAAG
MntH-4-F GATAGTAAGAGCTTGGACGAAGTC 225
MntH-4-R CCATGGCTTGTAGCAACATGGCAAT
MntH-5-F GCAGTCGGCTATATGGATCCCGGCAAC 246
MntH-5-R CATATCAGTCGCCATCATCGCGAC
MntR-F TATTTTCGAGCTTGGTGGTACGAA 218
MntR-R CTAAGAAATCTTCCCAGATTCGGTG
16S rRNA-F AGCAGCCGCGGTAATACGTAGGTG 227
16S rRNA-R GACCAGACAGCCGCCTTCGCCACAC


3. Results and discussion

3.1 Effects of manganese starvation on growth performance

Manganese is an essential element for cell growth and is involved in a variety of metabolic pathways. The growth of L. plantarum CCFM436 was inhibited by manganese starvation stress, as indicated by the lower OD values. Cell performance is shown in Fig. 1A. The final OD600 was 5.27 in MRS culture and 2.37 under manganese starvation stress. Similarly, μmax changes occurred with cell growth (see Fig. 1B). The μmax of the cells decreased from 0.310 to 0.256 h−1 under manganese starvation conditions. Moreover, the μmax peak time was significantly delayed, and the appearance time for μmax extended from 5.0 to 8.5 h. Growth was clearly limited, as observed in a previous report, in which the growth of L. plantarum NC8 was directly correlated with the amount of manganese ions in the medium.11 The highest OD600 was reached at microtiter scale at the highest manganese ion concentration tested. As manganese can participate as an enzyme cofactor in various biological processes, manganese starvation was harmful to the growth of L. plantarum CCFM436. However, slight cell growth was observed in complex media (e.g., MRS) without manganese addition. This may be caused by the yeast extract, beef extract or trypsin powder used containing trace amounts of manganese that could be used for cell growth.11
image file: c7ra00072c-f1.tif
Fig. 1 Effect of Mn starvation on growth kinetics of L. plantarum CCFM 436: (A) pH/biomass (OD600) and (B) μ.

3.2 Effects of manganese starvation stress on cell physiology

3.2.1 Cell morphology under manganese starvation stress. As the primary barrier, cell morphology is crucial in maintaining cell viability and metabolic function, particularly under environmental stress.18,20 Cell integrity is necessary to maintain the normal physiological function of cells and is essential for keeping intracellular biochemical reactions steady. As shown in Fig. 2A, compared with that of L. plantarum CCFM436 grown in normal MRS, the cell morphology of L. plantarum CCFM436 grown under manganese starvation had lower integrity. The cell membrane was also thinner. This cell shrinkage behavior might be beneficial for saving energy to better resist environmental stress.
image file: c7ra00072c-f2.tif
Fig. 2 Effect of Mn starvation on cell physiology of L. plantarum CCFM 436: (A) cell morphology and (B) cell membrane fatty acid composition. Fatty acids in colored bars, from bottom to top, are C12:0, C14:0, C16:0, C18:0, C16:1, C18:1, C18:1 (trans-1,3-octadecenoic acid), C18:2, C19-cyc and unidentified fatty acids, respectively.
3.2.2 Fatty acid composition of cell membrane under manganese starvation stress. As another main mechanical barrier, the cell membrane can defend against external stress by adjusting its fatty acid composition.21 The saturated fatty acids in L. plantarum CCFM436 were lauric (C12:0), myristic (C14:0), palmitic (C16:0) and stearic acids (C18:0), whereas the unsaturated acids were hexadecenoic (C16:1), oleic (C18:1), trans-13-octadecenoic (C18:1), linoleic (C18:2) and cyclopropane fatty acids (C19-cyc) (Fig. 2B). Unsaturated fatty acids dominated the fatty acid composition at all manganese levels. The proportion of unsaturated fatty acids increased from 69.45% to 77.02% under the manganese starvation conditions, which suggested that the higher proportion of unsaturated fatty acids in the cell membrane might contribute to adaption to the external environment. Furthermore, the ratio of saturated fatty acids to unsaturated fatty acids could indirectly reflect the mobility of the cell membrane. The unsaturated fatty acid content in the cell membrane of L. plantarum CCFM436 increased under manganese starvation conditions, causing cell membrane fluidity to increase. A higher proportion of unsaturated fatty acids is closely associated with certain types of metabolism (nutrient transport and exchange).22

Meanwhile, the cyclopropane fatty acid content was much higher under manganese starvation stress conditions (increased from 25.82% to 43.54%), whereas that of octadecenoic acid was significantly lower. High cyclopropane fatty acid contents have been reported as beneficial to maintaining better membrane compactness and resistance to harmful environmental factors.20,21

3.3 Intracellular amino acid composition and key functional groups under manganese starvation stress

3.3.1 Intracellular amino acid composition under manganese starvation stress. It has been suggested that amino acid metabolism plays an important role in regulating intracellular homeostasis, producing metabolic energy and enhancing the resistance of cells to environmental stress.23 Due to the lack of a TCA cycle, the main ATP production method in L. plantarum is amino acid metabolism. Contents of Asp, Glu, Arg, Lys, Thr and other amino acids were accumulated under manganese starvation stress (see Fig. 3A). Arg can be metabolized to produce ATP through the ADI pathway, while the decarboxylation reaction of Asp also provides energy for cellular growth.24 Glu, Asp and Thr could also be translated to other amino acids, such as ornithine, alanine and proline, through a series of transamination and other reactions.24 Arg and Lys have double amino groups, while Asp and Glu have double carboxyl groups. The increased contents of these amino acids demonstrated their close association with manganese adsorption of L. plantarum CCFM436 from another point of view, which was consistent with our previous reports. Meanwhile, the Cys and Ile contents decreased. Cys can be degraded into pyruvic acid and hydrogen sulfide under aerobic conditions. Ile, belonging to the branched chain amino acid (BCAA), can be swiftly consumed and provide energy to maintain cell metabolism under environmental stress.24
image file: c7ra00072c-f3.tif
Fig. 3 Effect of Mn starvation on intracellular amino acid and functional groups of L. plantarum CCFM 436: (A) intracellular amino acids and (B) FTIR spectra analysis.
3.3.2 Key functional groups under manganese starvation condition. FTIR spectroscopy was employed to analyze changes in functional groups on the cell surface under manganese starvation conditions (see Fig. 3B). The FTIR spectrum of cells from normal MRS medium displayed absorption peaks at 1038, 1220, 1536, 1634, 2915 and 3274 cm−1. Comparatively, those under manganese starvation stress afforded more complex peaks due to the increased number of derivatives in normal MRS. In the Mn-starvation medium, the main characteristic absorption peaks of strain CCFM436 were stable, with several peaks associated with specific functional groups (1038, 1536, 1634 and 3274 cm−1) having moved slightly. The most fluctuation was associated with carboxyl and amino groups, which was caused by manganese deficiency. Fluctuation of the absorption peaks at 1634 cm−1 (C[double bond, length as m-dash]O) and 3274 cm−1 (O[double bond, length as m-dash]C–OH) was due to the change in stretching vibration of the carboxyl group without manganese ion complexation or the ion exchange effect.25,26 Similarly, movement of the peaks at 1038 and 1536 cm−1 indicated N–H fluctuation and the generation of some different compounds when the functional amino group of the protein did not uptake manganese ions.26,27 These results were also consistent with the main chemical bond activations in the Mn bioadsorption process (such as C[double bond, length as m-dash]O stretching, N–H bending, and O–H stretching).27

3.4 Proteomic analysis and functional classification of identified proteins

3.4.1 Carbohydrate and nitrogen metabolism. To understand the protein expression of L. plantarum CCFM436 under manganese starvation stress, 2D electrophoresis was used to analyze the proteomics of whole cell proteins. The protein spots of L. plantarum CCFM436 were mainly located in the acidic and neutral regions. Protein profiles were analyzed by Quest software. The different protein spots were identified by MALDI-TOF/TOF/MS, and results are shown in Fig. 4 and Table 3. Among the 73 different protein spots, 19, 2, and 7 protein spots were involved in carbohydrate metabolism, amino acid and nitrogen metabolism, and nucleotide metabolism, respectively. Twenty-three protein spots participated in the transcription and translation process. Meanwhile, seven and five protein spots were associated with the transport system and stress response, respectively, and 10 protein spots were involved in other response processes.
image file: c7ra00072c-f4.tif
Fig. 4 Effect of Mn starvation on overall protein expression of L. plantarum CCFM 436 via proteome analysis.
Table 3 Identification of differentially expressed proteins with different levels of manganese in L. planturam CCFM436 by MALDI-TOF/TOF analysis
Spot Accession Protein name Theoretical MW (Da) MASCOT score No. matched Fold change 0/16
Carbon metabolism
2 gi|489735124 F0F1 ATP synthase subunit beta 50[thin space (1/6-em)]786 395 7(1) 1.63
3 gi|334880343 D-Lactate dehydrogenase (D-LDH) 37[thin space (1/6-em)]176 321 8(2) 3.86
6 gi|5823364 PepQ 41[thin space (1/6-em)]316 387 5(3) 0.84
7 gi|3043367 L-Lactate dehydrogenase 34[thin space (1/6-em)]256 90 2(0) 2.48
8 gi|334881845 Ribose-5-phosphate isomerase A 24[thin space (1/6-em)]617 217 4(1) 2.66
9 gi|334880734 NADH peroxidase 48[thin space (1/6-em)]443 560 9(6) 2.44
10 gi|489733775 Glyceraldehyde-3-phosphate dehydrogenase 36[thin space (1/6-em)]645 148 2(1) 1.83
12 gi|489733459 Fructose-bisphosphate aldolase 31[thin space (1/6-em)]060 851 9(7) 1.76
16 gi|334881745 Phosphoglycerate dehydrogenase 34[thin space (1/6-em)]635 82 1(1) 4.27
30 gi|334881069 Acetaldehyde dehydrogenase 49[thin space (1/6-em)]062 280 5(2) 3.20
34 gi|13940254 Acetate kinase 43[thin space (1/6-em)]870 364 6(4) 1.20
45 gi|489735713 Phosphoglyceromutase 26[thin space (1/6-em)]069 504 7(3) 0.34
47 gi|334882090 ATP synthase gamma chain 34[thin space (1/6-em)]586 225 3(2) 5.33
53 gi|300494598 Triose-phosphate isomerase 25[thin space (1/6-em)]858 268 3(2) 0.13
59 gi|489734188 Phosphoenolpyruvate-protein phosphotransferase 63[thin space (1/6-em)]271 67 1(1) 1.28
65 gi|334880497 Putative manganese-dependent inorganic pyrophosphatase 33[thin space (1/6-em)]645 563 7(3) 0.36
66 gi|489735164 Glucose-6-phosphate isomerase 49[thin space (1/6-em)]776 156 2(1) 2.47
67 gi|334880778 Ferredoxin-NADP reductase (FNR) (Fd-NADP + reductase) 36[thin space (1/6-em)]231 251 5(1) 2.05
68 gi|254046929 Putative phosphoketolase 90[thin space (1/6-em)]611 96 2(1) 0.30
[thin space (1/6-em)]
Nitrogen and amino acid metabolism
28 gi|334881598 Putative protein-tyrosine phosphatase 29[thin space (1/6-em)]819 522 7(4) 0.97
70 gi|334881735 Cysteine desulfurase 44[thin space (1/6-em)]929 114 2(0) 0.34
[thin space (1/6-em)]
Nucleotide metabolism
26 gi|334882300 Ribose-phosphate pyrophosphokinase 35[thin space (1/6-em)]709 368 6(2) 2.41
35 gi|334881576 Dihydroorotate dehydrogenase 31[thin space (1/6-em)]458 154 3(1) 0.49
44 gi|489733564 CTP synthase 59[thin space (1/6-em)]728 242 6(0) 1.08
48 gi|339638332 GMP reductase 35[thin space (1/6-em)]531 98 2(0) 0.09
50 gi|334881578 Carbamoyl-phosphate synthase pyrimidine-specific small chain 40[thin space (1/6-em)]289 366 8(1) 84.63
56 gi|544928213 Nucleoside-triphosphate diphosphatase 21[thin space (1/6-em)]831 142 2(2) 3.27
64 gi|334881512 Adenylosuccinate synthetase 47[thin space (1/6-em)]453 395 6(3) 0.53
[thin space (1/6-em)]
Transcription
4 gi|328813989 DNA-directed RNA polymerase alpha chain 29[thin space (1/6-em)]284 195 2(2) 2.43
25 gi|334881701 Primosomal protein DnaI 35[thin space (1/6-em)]367 158 2(1) 2.41
32 gi|489733165 DNA polymerase III subunit beta 41[thin space (1/6-em)]439 345 6(2) 1.69
41 gi|334881375 Chromosomal replication initiator protein dnaA 51[thin space (1/6-em)]385 185 2(1) 1.60
46 gi|334881483 Uncharacterized RNA methyltransferase lp_3226 53[thin space (1/6-em)]759 316 4(2) 1.76
61 gi|489733582 DEAD/DEAH box helicase 58[thin space (1/6-em)]875 146 3(1) 0.69
69 gi|640538996 DNA-directed RNA polymerase subunit beta' 135[thin space (1/6-em)]760 283 5(2) 0.26
[thin space (1/6-em)]
Translation
1 gi|339638926 Ribosomal protein S1 47[thin space (1/6-em)]212 404 5(3) 0.61
11 gi|334881031 L-2-Hydroxyisocaproate dehydrogenase 34[thin space (1/6-em)]810 280 4(2) 2.91
15 gi|489734520 Methionyl-tRNA formyltransferase 34[thin space (1/6-em)]473 37 6(3) 2.29
29 gi|334881840 Glutamyl-tRNA synthetase 57[thin space (1/6-em)]050 442 7(5) 2.59
38 gi|334881788 Ribosomal protein S30EA 21[thin space (1/6-em)]698 351 3(3) 0.38
42 gi|334882268 Elongation factor Tu (EF-Tu) 43[thin space (1/6-em)]378 138 2(1) 3.20
43 gi|334882897 Arginyl-tRNA synthetase 63[thin space (1/6-em)]003 71 1(1) 2.43
49 gi|489734874 30S ribosomal protein S2 30[thin space (1/6-em)]208 250 4(1) 0.15
51 gi|334881700 Threonyl-tRNA synthetase 73[thin space (1/6-em)]784 256 4(2) 4.76
52 gi|334882583 Aspartyl/glutamyl-tRNA (Asn/Gln) amidotransferase subunit B (Asp/Glu-ADT subunit B) 53[thin space (1/6-em)]315 259 5(2) 0.37
58 gi|334880391 Aspartyl-tRNA synthetase 68[thin space (1/6-em)]060 105 2(1) 1.24
54 gi|334881831 50S ribosomal protein L10 18[thin space (1/6-em)]212 495 6(3) 0.29
55 gi|334881832 50S ribosomal protein L1 24[thin space (1/6-em)]742 117 3(0) 0.18
57 gi|489733962 Elongation factor P 76[thin space (1/6-em)]909 162 3(2) 2.02
62 gi|489734874 30S ribosomal protein S2 30[thin space (1/6-em)]208 282 5(3) 0.25
63 gi|489733967 50S ribosomal protein L2 30[thin space (1/6-em)]145 342 6(3) 0.13
[thin space (1/6-em)]
Stress response
19 gi|489735106 Universal stress protein UspA 17[thin space (1/6-em)]728 445 4(4) 4.16
31 gi|334882861 ATP-dependent Clp protease, ATP-binding subunit ClpL 75[thin space (1/6-em)]929 224 3(2) 1.89
71 gi|1552551 Catabolite control protein A 30[thin space (1/6-em)]742 376 5(3) 0.64
72 gi|33312966 Heat shock protein 19[thin space (1/6-em)]729 165 2(2) 27.26
73 gi|334881995 UPF0082 protein lp_2253 26[thin space (1/6-em)]371 142 2(1) 1.75
[thin space (1/6-em)]
Transporter
14 gi|503121615 Glycine/betaine ABC transporter ATP-binding protein 44[thin space (1/6-em)]005 225 3(2) 2.14
22 gi|334882144 Multiple sugar ABC transporter, ATP-binding protein 41[thin space (1/6-em)]441 643 8(5) 3.13
24 gi|334882864 Oligopeptide ABC transporter, ATP-binding protein 36[thin space (1/6-em)]667 353 7(4) 1.59
33 gi|489733642 PTS mannose transporter subunit IIAB 35[thin space (1/6-em)]268 573 6(6) 2.19
36 gi|334883019 Glutamine ABC transporter, ATP-binding protein 26[thin space (1/6-em)]811 189 2(2) 6.35
37 gi|334882691 Metal-dependent regulator 23[thin space (1/6-em)]873 152 4(0) 0.50
60 gi|334882868 Oligopeptide ABC transporter, substrate binding protein 60[thin space (1/6-em)]321 66 1(1) 6.01
[thin space (1/6-em)]
Others
5 gi|334880815 Lipoate-protein ligase 1201 262 4(2) 27.26
13 gi|334882009 2,3,4,5-Tetrahydropyridine-2,6-dicarboxylate N-acetyltransferase 24[thin space (1/6-em)]552 282 3(2) 0.93
17 gi|334882348 Metallo-beta-lactamase 27[thin space (1/6-em)]604 109 2(1) 2.07
18 gi|334881906 Putative uncharacterized protein 20[thin space (1/6-em)]742 105 1(1) 6.94
20 gi|489733483 Hypothetical protein 18[thin space (1/6-em)]207 246 3(1) 2.19
21 gi|300493552 Ribose-phosphate diphosphokinase 35[thin space (1/6-em)]551     2.06
23 gi|334880640 Phosphate acyltransferase 37[thin space (1/6-em)]400 293 4(2) 6.75
27 gi|334880612 3-Oxoacyl-[acyl-carrier-protein] synthase 3 protein 2 35[thin space (1/6-em)]285 206 3(2) 2.23
40 gi|334882817 Poly(glycerol-phosphate) alpha-glucosyltransferase 57[thin space (1/6-em)]597 198 4(2) 2.39
39 gi|334882270 Putative metallo-beta-lactamase superfamily protein 63[thin space (1/6-em)]946 195 4(0) 1.77


Cellular metabolism and flow direction were essential to the growth of L. plantarum CCFM436. Nineteen proteins were involved in carbon metabolism under manganese starvation conditions, including glyceraldehyde-3-phosphate dehydrogenase (spot 10), fructose-2-phosphate aldolase (spot 12), phosphoglycerate (spot 16), glucose-6-phosphate isomerase (spot 66) and other proteins involved in the glycolysis pathway. As a type of facultative anaerobic bacteria, the glycolysis pathway is the main energy source of L. plantarum without a TCA cycle.28,29 Meanwhile, tyrosine phosphatase and cysteine desulfurase participated in nitrogen metabolism, and CTP synthetase, dihydrolactate dehydrogenase and phosphate ribose pyrophosphate kinase were associated with nucleotide metabolism.

In terms of energy metabolism, the expression of some proteins involved in the glycolytic pathway for ATP synthesis were upregulated. For example, expression of glycosyl-bisphosphate aldolase (spot 12) and glyceraldehyde-3-phosphate dehydrogenase (spot 10) were upregulated 1.83-fold and 1.76-fold, respectively. The glycolytic pathway could be promoted through upregulation of these proteins. The expression of F0F1 ATP synthase beta (spot 2) and ATP synthase γ (spot 47) subunits, which are directly involved in the synthesis of ATP, were also upregulated 1.63-fold and 5.33-fold, respectively. Furthermore, the expression of ribose-5-phosphate isomerase (spot 8), which is involved in the pentose phosphate pathway, was also significantly increased, and its expression was upregulated 2.6-fold, indicating that pentose phosphate pathway activity was elevated under manganese starvation and provided more reduction power (NADPH) for the growth of L. plantarum CCFM436.

During amino acid metabolism, cysteine desulfurase (spot 70) catalyzed the desulfurization of cysteine to form alanine and enzyme-sulfide intermediates. The damage to superoxide ions in the cells could not be removed in time under manganese starvation conditions. Therefore, the expression of cysteine desulfurase was decreased, thereby increasing intracellular free cysteine content and regulating redox balance in the cells.8

3.4.2 Transcription and translation. Seventeen proteins participated in the transcription and translation process under manganese starvation conditions. The expression of chromosomal replication initiator protein DnaA (spot 41), DNA polymerase III subunit beta (spot 32) and DNA-directed RNA polymerase alpha chain (spot 4) were upregulated. Due to the functional association in tRNA synthesis, some proteins involved in aminoacyl-tRNA biosynthesis, such as glutamyl-tRNA synthetase (spot 29), arginyl-tRNA synthetase (spot 43) and aspartyl-tRNA synthetase (spot 58), were upregulated, indicating that L. plantarum improved many biological pathways by activating the translational level. However, the expression of aspartyl/glutamyl-tRNA (Asn/Gln) amidotransferase subunit B (spot 52) was downregulated. The AtMTP family of genes encodes proteins of the cation diffusion facilitator (CDF) family, which has several members with roles in manganese tolerance and is implicated in the transport of manganese as a tolerance mechanism.30 Furthermore, the expression of ribosomal proteins (spots 1, 54, 55, 62 and 63) was downregulated, whereas expression of EF-Tu (spot 42) and EF-P (spot 57) was upregulated, suggesting a decrease in protein synthesis. These elongation factors are mostly responsible for escorting aminoacyl tRNAs to the ribosome as they proceed along mRNA.1,8 Nevertheless, these factors seem to have a chaperone-like function in the protein-folding process. The interaction between the ribosome and elongation factors complexed to aminoacyl tRNA is essential for controlling translational accuracy.8 EF-Tu has been reported as a factor in the adhesion of L. plantarum to manganese.29 The expression of ribosomal proteins and elongation factors changed, along with the possible enhancement of translation accuracy and protein folding.
3.4.3 Molecular chaperone and stress response metabolism. Several stress response proteins were identified through MS analysis, such as heat shock protein (HSP) (spot 72), universal stress protein UspA (spot 19), ATP-dependent Clp protease (spot 31) and UPF0082 protein (spot 73). As the group of functional proteins responsible for protein folding, HSPs are closely related to environmental stresses, such as heat and heavy-metal stress.8,30 HSP (spot 72) and universal stress protein UspA (spot 19) were upregulated 26.26-fold and 4.16-fold, respectively, under manganese starvation conditions. Meanwhile, the expression of ATP-dependent Clp protease, which was involved in molecular chaperoning and corrected protein folding, or cleaved/degraded misfolded proteins, was upregulated 2.89-fold. This molecular chaperone and stress response protein could be involved in the repair or degradation of misfolded proteins to improve protein folding accuracy.1,8

3.5 Transcription analysis of manganese transporter genes using qRT-PCR

As the key regulator and defender in controlling intracellular steady-state manganese balance, the expression of manganese transporters is closely related to changes in manganese content.31,32 However, identifying manganese transporters located in the membrane by 2D electrophoresis was difficult, so qRT-PCR analysis was performed instead (Fig. 5). Manganese uptake occurred through different types of cation transporters, with an active manganese ion transport system and five manganese transporters (MntH 1–5) identified in our previous study. The expression and regulation of manganese transporters were significantly different under manganese starvation stress. The transcription level of the manganese transporter regulator (MntR) was downregulated 0.69-fold, while those of manganese transporters (MntH 1–5) were upregulated. The transcription level of MntH 2 was upregulated more than 25-fold. The negatively regulation model between MntR and MntH was consistent with transcription levels changes in MntH 1–5 under 960 mg L−1 manganese stress (data not published). MntR is a transcriptional regulator that binds two Mn atoms in manganese-replete cells and then bonds with MntH promoters to form MntR:Mn2. Therefore, the synthesis of manganese transporters can be downregulated. As the primary signal, Mn2+ can regulate the expression of Mn2+ transporters via Mn2+-dependent repressor MntR.31 The results suggest that MntH 1–5 were responsible for intracellular importation of manganese into L. plantarum, which acted as an importer under manganese starvation stress. It was necessary to acquire manganese to maintain normal L. plantarum metabolism. Mn starvation-induced metabolic changes to key elements in L. plantarum CCFM 436 are summarized in Fig. 6.
image file: c7ra00072c-f5.tif
Fig. 5 Effect of Mn starvation on key related Mn transporters of L. plantarum CCFM 436.

image file: c7ra00072c-f6.tif
Fig. 6 Overview of the Mn starvation-induced metabolic changes in key elements of L. plantarum CCFM 436.

4. Conclusion

By analyzing cell physiology, proteomics and transporters, the response mechanisms of L. plantarum to Mn starvation stress were investigated. The results showed that μmax decreased from 0.310 to 0.256 h−1, TEM analysis indicated thinner cell morphology, and GC-MS analysis showed a higher proportion of unsaturated fatty acids and cyclopropane fatty acids, which improved the membrane fluidity and compactness. HPLC analysis of intracellular free amino acids indicated that intracellular Asp, Glu and Arg contents, which are closely related to energy metabolism, were increased. FTIR analysis showed that amino and carboxyl functional groups were significantly affected by the Mn starvation conditions. With comparative two-dimensional proteomic analysis, 73 proteins differentially expressed proteins involved in carbohydrate, amino acid and transcription/translation metabolisms, and stress response was identified under the Mn starvation conditions. Moreover, the potential Mn importer role of MntH 1–5 (negatively regulated by MntR) under Mn starvation stress was demonstrated using qRT-PCR analysis. The coordinated metabolism regulation model provides new insights into the response mechanism of probiotics under similar metal-starvation stress conditions.

Acknowledgements

This work was supported by State Key Program of National Natural Science Foundation of China (No. 31530056), the Science and Nature Foundation of Jiangsu Province (No. BK 20160169), the Jiangsu Province Postdoctoral Science Foundation Funded Project (No. 1601183C), the China Postdoctoral Science Foundation Funded Project (No. 2015M581727), the Program of Introducing Talents of Discipline to Universities (B07029), the Program for Changjiang Scholars and Innovative Research Team in University (IRT1249), and the Program of Collaborative innovation center of food safety and quality control in Jiangsu Province.

Notes and references

  1. S. I. Hsieh, M. Castruita, D. Malasarn, E. Urzica, J. Erde, M. D. Page, H. Yamasaki, D. Casero, M. Pellegrini, S. S. Merchant and J. A. Loo, Mol. Cell. Proteomics, 2013, 12, 65–86 Search PubMed.
  2. B. A. Eijkelkamp, C. A. McDevitt and T. Kitten, BioMetals, 2015, 28, 491–508 CrossRef CAS PubMed.
  3. X. Zhang, H. H. Yang and Z. J. Cui, RSC Adv., 2016, 6, 27963–27968 RSC.
  4. S. B. Schmidt, P. E. Jensen and S. Husted, Trends Plant Sci., 2016, 21, 622–632 CrossRef CAS PubMed.
  5. M. D. Angelis and M. Gobbetti, Appl. Microbiol. Biotechnol., 1999, 51, 358–363 CrossRef PubMed.
  6. E. J. Gudiña, E. C. Fernandes, J. A. Teixeira and L. R. Rodrigues, RSC Adv., 2015, 5, 90960–90968 RSC.
  7. J. N. Bhakta, K. Ohnishi, Y. Munekage, K. Iwasaki and M. Q. Wei, J. Appl. Microbiol., 2012, 112, 1193–1206 CrossRef CAS PubMed.
  8. H. An, F. P. Douillard, G. Wang, Z. Zhai, J. Yang, S. Song, J. Cui, F. Ren, Y. Luo, B. Zhang and Y. Hao, Mol. Cell. Proteomics, 2014, 13, 2558–2572 CAS.
  9. S. H. Ye, M. P. Zhang, H. Yang, H. Wang, S. Xiao, Y. Liu and J. H. Wang, Carbohydr. Polym., 2014, 30, 50–56 Search PubMed.
  10. S. Das, H. R. Dash and J. Chakraborty, Appl. Microbiol. Biotechnol., 2016, 100, 2967–2984 CrossRef CAS PubMed.
  11. N. Böhmer, S. König and L. Fischer, FEMS Microbiol. Lett., 2013, 342, 37–44 CrossRef PubMed.
  12. L. C. Lew, M. T. Liong and C. Y. Gan, J. Appl. Microbiol., 2012, 114, 526–535 CrossRef PubMed.
  13. L. Karaffa, R. Díaz, B. Papp, E. Fekete, E. Sándor and C. P. Kubicek, Appl. Microbiol. Biotechnol., 2015, 99, 7937–7944 CrossRef CAS PubMed.
  14. S. I. Hsieh, M. Castruita, D. Malasarn, E. Urzica, J. Erde, M. D. Page, H. Yamasaki, D. Casero, M. Pellegrini and S. S. Merchant, Mol. Cell. Proteomics, 2013, 12, 65–86 Search PubMed.
  15. Q. X. Zhai, F. W. Tian, G. Wang, J. X. Zhao, X. M. Liu, K. Cross, H. Zhang, A. Narbad and W. Chen, RSC Adv., 2016, 6, 5990–9997 RSC.
  16. J. L. Xing, G. Wang, Z. N. Gu, X. M. Liu, Q. X. Zhang, J. X. Zhao, H. Zhang, Y. Q. Chen and W. Chen, RSC Adv., 2015, 5, 37626–37634 RSC.
  17. J. N. Bhakta, K. Ohnishi, Y. Munekage, K. Iwasaki and M. Q. Wei, J. Appl. Microbiol., 2012, 112, 1193–1206 CrossRef CAS PubMed.
  18. E. Dertli, M. J. Mayer and A. Narbad, BMC Microbiol., 2015, 15, 8–16 CrossRef PubMed.
  19. Y. J. Tong, G. Wang, Q. X. Zhang, F. W. Tian, X. M. Liu, J. X. Zhao, H. Zhang and W. Chen, RSC Adv., 2016, 6, 102804–102813 RSC.
  20. E. I. Urzica, A. Vieler, A. Hong-Hermesdorf, M. D. Page, D. Casero, S. D. Gallaher, J. Kropat, M. Pellegrini, C. Benning and S. S. Merchant, J. Biol. Chem., 2013, 288, 30246–30258 CrossRef CAS PubMed.
  21. Y. M. Zhang and C. O. Rock, Nat. Rev. Microbiol., 2008, 6, 222–233 CrossRef CAS PubMed.
  22. H. J. Hwang, S. P. Choi, S. Y. Lee, J. I. Choi, S. J. Han and P. C. Lee, J. Biotechnol., 2015, 193, 130–133 CrossRef CAS PubMed.
  23. K. Senouci-Rezkallah, P. Schmitt and M. P. Jobin, Food Microbiol., 2011, 28, 364–372 CrossRef CAS PubMed.
  24. P. Newsholme, L. Stenson, M. Sulvucci, R. Sumayao and M. Krause, Compr. Biotechnol., 2011, 27, 3–14 Search PubMed.
  25. K. M. H. Mata, O. M. Amaya, M. T. C. Barragan, F. J. A. Tapia and E. A. Felix, Int. J. Photoenergy, 2013, 578729 Search PubMed.
  26. A. Hernandez, D. Guzman and H. Garcia, J. Appl. Microbiol., 2009, 107, 395–403 CrossRef PubMed.
  27. M. Feng, X. Chen, C. Li, R. Nurgul and M. Dong, J. Food Sci., 2012, 77, T111–T117 CrossRef CAS PubMed.
  28. M. F. Mazzeo, R. Lippolis, A. Sorrentino, S. Liberti, F. Fragnito and R. A. Siciliano, PLoS One, 2015, 10, e0142376–0142389 Search PubMed.
  29. E. Vera Pingitore, A. Pessione, C. Fontana, R. Mazzoli and E. Pessione, Int. J. Food Microbiol., 2016, 238, 96–102 CrossRef CAS PubMed.
  30. E. Delhaize, B. D. Gruber, J. K. Pittman, R. G. White, H. Leung, Y. Miao, L. Jiang, P. R. Ryan and A. E. Richardson, Plant J., 2007, 51, 198–210 CrossRef CAS PubMed.
  31. T. E. Gunter, B. Gerstner, K. K. Gunter, J. Malecki, R. Gelein, W. M. Valentine, M. Aschner and D. I. Yule, Neurotoxicology, 2013, 34, 118–127 CrossRef CAS PubMed.
  32. Z. Ma, F. E. Jacobsen and D. P. Giedroc, Chem. Rev., 2009, 109, 4644–4681 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2017