Journal of Applied Microbiology ISSN 1364-5072

ORIGINAL ARTICLE

Assessment of stress tolerance acquisition in the heattolerant derivative strains of Bifidobacterium animalis subsp. lactis BB-12 and Lactobacillus rhamnosus GG nchez3,4, M. Gueimonde1,3 and S. Salminen1 J. Aakko1,2,3, B. Sa 1 2 3 4

Functional Foods Forum, University of Turku, Turku, Finland Food Chemistry and Food Development, Department of Biochemistry, University of Turku, Turku, Finland Department of Microbiology and Biochemistry of Dairy Products, IPLA-CSIC, Villaviciosa, Asturias, Spain Nutrition and Bromatology Group, Department of Analytical and Food Chemistry, Food Science and Technology Faculty, University of Vigo – Ourense Campus, Ourense, Spain

Keywords 2D-DIGE, Bifidobacterium animalis subsp. lactis BB-12, heat shock, Lactobacillus rhamnosus GG, probiotics, proteomics. Correspondence Juhani Aakko, Functional Foods Forum, University of Turku, It€ ainen Pitk€akatu 4 A, 5. krs, 20014 Turku, Finland. E-mail: [email protected] 2013/2186: received 31 October 2013, revised 2 April 2014 and accepted 3 April 2014 doi:10.1111/jam.12520

Abstract Aims: The purpose of this study was to investigate the heat-shock response at molecular level in Lactobacillus rhamnosus GG, Bifidobacterium animalis subsp. lactis BB-12 and their heat-tolerant derivatives and to characterize the changes that make the derivatives more robust in terms of heat stress. Methods and results: The study strains were exposed for 2 h to a heat-shock treatment, Bif. animalis subsp. lactis BB-12 and its derivative at 50°C and the Lact. rhamnosus GG and its derivative at 60°C. Protein synthesis before and after heat shock was examined using proteomics and RT-qPCR. The analysis revealed that the regulation of seven proteins in both strain pairs was modified as a response to heat or between the original and the derivative strain. The comparison of wild-type strains and the heat-tolerant derivatives suggests that the acquisition of heat tolerance in the Bif. animalis subsp. lactis BB-12 derivative is due to a slightly increased constitutive level of chaperones, while in Lact. rhamnosus GG derivative, the main reason seems to be a higher ability to induce the production of chaperones. Conclusions: This study revealed possible markers of heat tolerance in B. lactis and Lact. rhamnosus strains. Significance and impact of study: This study increases our knowledge on how Lactobacillus and Bifidobacterium strains may acquire heat tolerance. These findings may be useful for improving the heat tolerance of existing probiotic strains as well as screening new heat-tolerant strains.

Introduction Probiotics are defined as ‘live micro-organisms which when administered in adequate amounts, confer a health benefit on the host’ (FAO/WHO 2002). At present, lactobacilli and bifidobacteria are the most common bacteria used as probiotics, and their properties are strain dependent and strain specific. To exert their beneficial effects, probiotic bacteria must retain the characteristics and properties originally used for their selection (Tuomola et al. 2001).

During different manufacturing and storage stages, probiotic bacteria have to respond rapidly to stress in order to survive. Heat is among the hardest stress conditions that may be present during manufacturing of different food products, such as bread and chocolate, in which probiotics could be applicable. The poor heat tolerance of probiotics has precluded their inclusion in such products and limited their application to fermented milks and similar products where heat stress is not present during manufacturing. Exposure to high temperatures leads to denaturation of proteins and their subsequent aggregation. Heat also

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

239

Heat adaptation and response in LGG and BB-12

BB-12* WT† and BB-12 HS‡

J. Aakko et al.

LGG§ WT and LGG HS

Heat-treatment, 50°C for B. lactis and 60°C for L. rhamnosus, 0 and 2 h

Protein extraction

RNA extraction

2-D electrophoresis

RT-qPCR



Protein expression fold change in WT and HS between time points, 0 h vs 2 h



Gene expression change between time points in WT and HS strains, 0 h vs 2 h



Direct comparison of WT and HS strains at baseline and after heat treatment, WT 0 h vs HS 0 h,



Comparison of wild type and derivative strains, WT 0 h vs HS 0 h and WT 2 h vs HS 2 h

destabilizes macromolecules such as ribosomes and RNA, and alterations of membrane fluidity have been reported as well (van de Guchte et al. 2002). The mechanisms underlying heat-shock adaptation have been described earlier in different bacterial species, most notably in Escherichia coli and Bacillus subtilis (Richmond et al. 1999; Helmann et al. 2001; Stewart et al. 2002; Weiner et al. 2003; De Angelis and Gobbetti 2004; Gao et al. 2004; Barreiro et al. 2005; Rezzonico et al. 2007). In this study, we used two-dimensional differential ingel electrophoresis (2D-DIGE) coupled with mass spectrometry, together with reverse transcriptase-quantitative PCR (RT-qPCR), to assess the heat response in the two most commonly used probiotic strains, Bifidobacterium animalis subsp. lactis BB-12 and Lactobacillus rhamnosus GG (ATCC 53013), and to compare them with their previously obtained heat-tolerant derivatives (BB-12 HS and LGG HS) (du Toit et al. 2013). Our aim was to determine the molecular basis of the heat-stress response in these strains and to characterize the changes that make the derivative strains more robust in terms of heat stress.

Figure 1 A schematic workflow of the study. *BB-12 = Bifidobacterium animalis subsp. lactis BB-12. †WT = wild-type strain. ‡ HS = heat-tolerant derivative strain. § LGG = Lactobacillus rhamnosus GG.

Systems) supplemented with 005% L-cysteine (w/v) (Sigma-Aldrich, St. Louis, MO) (MRSC) in an anaerobic cabinet (MG500; Don Whitley Scientific, Yorkshire, UK) with an atmosphere of 10% CO2, 10% H2 and 80% N2 at 37°C. 05 ml of these precultures was inoculated to 50 ml of MRSC medium and grown overnight at 37°C in the same conditions. After 16 h, biomass was harvested. The cultures were centrifuged at 15 000 g and resuspended in 50 ml of fresh medium. The heat treatment was performed by incubating the Bif. animalis subsp. lactis BB12 and BB-12 HS cultures at 50°C and those of Lact. rhamnosus GG at 60°C for 2 h (Fig. 1). Samples were taken before and after the heat treatment; the cultures were centrifuged at 15 000 g and washed twice with PBS, and then, the bacterial pellets were frozen at 20°C until protein extraction. For RT-qPCR, after the heat treatment and washing of the cultures, 100 ll of the bacterial suspension was diluted in 400 ll of PBS buffer and 1000 ll of RNA protect bacteria reagent (Qiagen, Venlo, Netherlands) was added to stabilize the RNA according to the manufacturer’s instructions. The stabilized pellets were stored at 80°C until RNA extraction.

Materials and methods Protein extraction Bacterial strains and growth Bacterial strains used in this study were Bif. animalis subsp. lactis BB-12 and its heat-tolerant derivative (BB-12 HS) and Lact. rhamnosus GG (ATCC 53013) and its heat-tolerant derivative (LGG HS) (du Toit et al. 2013). Strains were routinely grown in de Man-Rogosa-Sharpe broth (De Man et al. 1960) (MRS; BD Diagnostic 240

The cells were broken by ultrasonic treatment (Vibra-cell, model VC600, 20 kHz; Sonics & Materials, Inc., Newtown, CT) by sonicating the cells three times for 1 min and keeping them on ice for 1 min in between the sonications. Next, unbroken cells and cell debris were removed by centrifugation (15 000 g at 4°C, 5 min). The proteins were precipitated by methanol/chloroform

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

J. Aakko et al.

(3 : 1, v/v) according to the method described by Wessel and Flugge (1984), and the protein concentration of the extract was determined using Pierce BCA Protein Assay Kit (Thermo Fisher Scientific, Waltham, MA) according to the manufacturer’s instructions. The protein extracts were frozen at 20°C until further use. 2D Electrophoresis The 2D electrophoresis was performed with two different dyeing techniques. First, the electrophoresis was visualized using Gelcode Blue Safe Protein Stain (Thermo scientific) to obtain reference maps of the proteins in the different extracts. Additionally, these gels were used for spot picking and mass spectrometry analysis. The same protein extracts were also analysed with Amersham CyDye DIGE Fluors (minimal dyes) for Ettan DIGE (GE Healthcare). For the 2D electrophoresis, 500 lg of protein was used for the reference map gels and 50 lg for the gels on which the samples were labelled with fluorescent dyes. For the reference map gels, the protein samples were solubilized in 450 ll Destreak Rehydration Solution (GE Healthcare, Little Chalfont, UK) supplemented with 05% IPG buffer (pH 4–7; GE Healthcare). The solutions were then used to rehydrate Immobiline Drystrips (pH 4–7, 24 cm; GE Healthcare) for 12 h at constant voltage of 50 V at 20°C using an Ettan IPGphor device (Amersham Biosciences, Amersham, UK) and focused at 60 000 Vh. The focused strips were then equilibrated for 15 min in a buffer (1 mol l 1 Tris-HCl, pH 68, containing 6 mol l 1 urea, 30% (v/v) glycerol and 1% (w/v) sodium dodecyl sulphate) supplemented with 083% (w/v) dithiothreitol in the first equilibration step and with 75% (w/v) iodoacetamide in the second one. The second-dimension separation was performed by sodium dodecyl sulphatepolyacrylamide gel electrophoresis in 125% polyacrylamide gels and carried out with Ettan DALTsix electrophoresis unit (Amersham Biosciences) at a constant current of 16 mA per gel overnight at 4°C. After the electrophoresis, the gels were stained with Gelcode Blue Safe Protein Stain (Thermo scientific) according to the manufacturer’s instructions and finally, the gels were scanned with Imagescanner using LABSCAN software (Amersham Biosciences). 2D-DIGE with Amersham CyDye DIGE Fluors (minimal dyes) for Ettan DIGE (6905-6919 GE Healthcare) As mentioned earlier, protein spots were also visualized with fluorescent dyes. After the precipitation, the samples were solubilized in 20 ll of lysis buffer (30 mmol l 1 Tris, 7 mol l 1 urea, 2 mol l 1 thiourea and 4% (w/v)

Heat adaptation and response in LGG and BB-12

CHAPS at pH 85). Each sample was labelled with different Amersham CyDye DIGE Fluors (minimal dyes) for Ettan DIGE (GE Healthcare). Briefly, 333 pmol of dye was used for 50 lg of protein, and the samples were incubated on ice for 30 min in the dark. After the incubation, 1 ll of 10 mmol l 1 lysine was added to stop the reaction and the samples were incubated on ice in the dark for 10 min. Twenty microlitre of 29 sample buffer (8 mol l 1 urea, 130 mmol l 1 DTT, 4% (w/v) CHAPS, 2% (v/v) IPG buffer, pH 4–7; GE Healthcare) was added to the samples, and again, they were incubated on ice in the dark for 10 min. Samples for the same gels were pooled together, and 330 ll of rehydration buffer (8 mol l 1 urea, 4% (w/v) CHAPS, 1% (v/v) IPG buffer, pH 4–7 and 13 mmol l 1 DTT) was added to the pooled samples. Immobiline Drystrips (pH 4–7, 24 cm; GE Healthcare) were rehydrated with the protein solutions, and both first-dimensional isoelectric focusing and second-dimension separation were carried out as described earlier for the protein reference maps. Ultimately, the gels were scanned in a Typhoon 9400 scanner (GE Healthcare) at a resolution of 100 lm. Spots showing clear difference between the time points and strains were chosen for further analysis. Spot detection and normalized volume quantization were carried out with IMAGEMASTER 2D PLATINUM (version 6.0 DIGE; GE Healthcare). The effect of heat shock on protein production was analysed by spot matching across gels, normalized volume ratio computing (0 vs 2 h). Spots that displayed a fold change of at least 15 between the time points or between the original and the derivative strain were considered significant (Fig. 1). Differences between the spot-normalized volumes were analysed statistically with the Student’s t-test at a significance level of P < 005. Spots showing statistical differences among conditions were excised from the 2D-PAGE gels visualized with Gelcode stain and sent to ‘Unidad de Prote omica- Fundaci on Centro Nacional de Investigaciones Cardiovasculares Carlos III’ (Madrid, Spain) for in-gel trypsin digestion and identification by matrix-assisted laser desorption/ionization coupled to time of flight mass spectrometry (MALDI-TOF) analysis. The in-gel trypsination, mass spectrometry analysis, was performed as previously described by Dominguez et al. (2009). BIOTOOLS 3.0 program (Bruker Daltonik, Bremen, Germany) was used to combine the MALDI-MS and MS/MS data and search a nonredundant protein database (NCBInr; c. 107 entries; National Center for Biotechnology Information, Bethesda, USA; or SwissProt; c. 5 9 105 entries; Swiss Institute for Bioinformatics, Switzerland) using the Mascot software (Matrix Science, London, UK). Other relevant search parameters were set as

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

241

Heat adaptation and response in LGG and BB-12

J. Aakko et al.

Table 1 Specific primers used for RT-qPCR analyses Spot

Protein

Locus tag

Gene

BL1

GroEL

BIF_00400

BL2

GroES

BIF_00675

BL4

Thioredoxin peroxidase

BIF_00644

BL5

Elongation factor G

BIF_00045

BL6

Transketolase

BIF_0093

BL7

Elongation factor Tu

BIF_01972

LR1

Chaperonin GroEL

LRHM_2152

LR2

Molecular chaperone, DnaK

LGG_01604

dnaK

LR3

LGG_00983

glmS

LR4

Glucosamine–fructose-6-phosphate aminotransferase, isomerizing L-lactate dehydrogenase

LGG_02523

ldh

LR5

Phosphocarrier protein Hpr

LGG_01821

ptsH

LR6

Trigger factor

LGG_01351

tig

LR7

Elongation factor G

LGG_02493

fusA

follows: enzyme, trypsin; fixed modifications, carbamidomethyl (C); allow up to one missed cleavages; peptide tolerance 20 ppm; MS/MS tolerance 05 Da. Verification and quantification of gene expression by reverse transcriptase-quantitative PCR Primers for different genes identified as differentially expressed in the proteomic analysis were designed using PRIMER EXPRESS 2.0 software (Applied biosystems), Table 1. The RNA was isolated from the earlier-obtained culture samples with the Rneasy mini kit (Qiagen) according to the manufacturer’s instructions. RNase-free DNase set (Qiagen) was used for DNA removal, and the RNA was quantified at 260 and 280 nm with Biotek Take 3 apparatus. One microgram of RNA was then used for cDNA synthesis, which was performed using the High-Capacity cDNA Archive kit (Applied biosystems) following the provided protocol. The cDNA samples were stored at 80°C for further analyses. The quantitative PCR studies were carried out on an ABI Prism 7500 Fast Real-Time PCR System (Applied Biosystems) using SYBR Green master mix (Applied Biosystems). Each reaction mixture (25 ll) composed of SYBR Green master mix, 02 lmol l 1 of primers and 242

Primer

Sequence

Amplicon length (bp)

BL1F BL1R BL2F BL2R BL4F BL4R BL5F BL5R BL6F BL6R BL7F BL7R LR1F LR1R LR2F LR2R LR3F LR3R LR4F LR4R LR5F LR5R LR6F LR6R LR7F LR7R

TCGTCACCAATGCGGAAGAC TGCTCGAGGTGAGCAGAATGT AGGTTTCCTTGCCGTCCTTCT GTTCACCGTGCACAAGCTGTAC AGGTTTCCTTGCCGTCCTTCT GTTCACCGTGCACAAGCTGTAC CGGCATCATGGCACACATC CTTGTAGTTCTTACCGGTGTAGAACAGA TCTTTCTGGCGGTCATGCA CCAGCGTCAGGCCATAACC TCCAAGGTTCTGCACGATGAG GAGCGGCATCGATCTGGTT CCAAAGCCAGGTGCTTTAACG CTGCCGACCTTGGTTCTGAA ACAAGCCGCTGGAACTCTTG ATCGGCAAAGGACCTTGGTACT GACGCTAACCGCTGATGCA CGGCTTGCGATCTTCTACCTT TCGTGACGCCGCTTATGAA GGGCAGTTGCGATACCATAGAA ACTTGCTTGCTGCCTGTACCA TTGCAGAAACCGGGATCCA TGACCGTCCCTGGATTTCG GGCATCACCATACATCCGTTTG AGTCCGGTGGTAAAGGTCAATATG GCCCTTGCCTTCTTCGTTTG

66 62 62 79 71 70 71 65 64 65 60 68 68

1 ll of cDNA. Thermal cycling consisted of an initial cycle of 10 min at 95°C followed by 35 cycles of 15 s at 95°C and 1 min at 60°C. Expression levels were determined by relative quantification using the DDCt method, as described by Livak and Schmittgen (2001), in which the expression level in the control culture (time zero or original strain culture, depending on the comparison) is arbitrarily set to 1, and the expression levels in the other samples are calculated relative to that control. The comparisons performed are illustrated in Figure 1. The 16S rRNA was used as an endogenous control, and the samples were analysed in duplicate in two independent PCR runs. Results Several changes were observed when a proteomic approach was used to detect proteins potentially involved in the heat adaptation and response of Bif. animalis subsp. lactis BB-12, Lact. rhamnosus GG and their heat-tolerant derivative strains (Fig. 2). In addition to characterizing the heat response in general, our goal was to define how the heat-tolerant derivatives of the studied probiotic strains have acquired their heat tolerance. To this end, the protein expression profiles of these strains and their

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

J. Aakko et al.

Heat adaptation and response in LGG and BB-12

BB-12 MM (kDa)

LGG pI

4·0 BL6

BL5

pI

7·0 4·0 MM (kDa)

LR7 LR1

7·0 LR3

LR6

Figure 2 Representative 2D gels of the Bif. animalis subsp. lactis BB-12 and Lact. rhamnosus GG strains. Proteins spots identified in the mass spectrometry analysis are labelled. BB-12, Bifidobacterium animalis subsp. lactis BB-12; LGG, Lactobacillus rhamnosus GG; MM, molecular mass; pI, isoelectric point.

LR2

BL1 BL7

LR4

BL4

LR5

BL2 BL3

Table 2 Proteins identified in mass spectrometry analysis and regulation of the proteins after the 2 h heat treatment according to proteomic studies. Ratios represent increases/decreases of normalized spots volumes after 2 h of heat-shock exposure in the wild-type (WT) and heat-shockresistant (HS) strains (0 < ratio < 1 ? protein downregulation; 1 < ratio ? protein upregulation) Fold difference 2 h/0 h

Strain

Spot

Protein

BB-12

BL1 BL2 BL3 BL4

GroEL GroES Unidentified Thioredoxin peroxidase Unidentified Transketolase

BL5 BL6

BL7 LGG

LR1 LR2 LR3

LR4

Elongation factor Tu Chaperonin GroEL Molecular chaperone DnaK Glucosamine– fructose6-phosphate aminotransferase, isomerizing L-lactate dehydrogenase

LR5

Phosphocarrier protein Hpr

LR6 LR7

Trigger factor Elongation factor G

MM, theoretical molecular LGG = Lact. rhamnosus GG.

mass

2D-DIGE

RT-qPCR

MM

WT

HS

WT

gi|14719389 gi|51094292

56 456 7043

189 354

178 220

596  358 104  091

355  084 065  083

43

gi|183601582

18 567

588

244

073  038

081  106

127

27

gi|289178611

80 150

133

063

085  027

020  003

541

31

gi|183602587

44 281

081

035

308  078

057  041

chaperone chaperone

419 126

37 21

gi|259650517 gi|199597089

55 809 67 179

456 126

441 074

010  001 023  001

747  336 1136  939

glutamate metabolism

570

52

gi|258507978

65 788

068

045

065  025

268  271

carbohydrate metabolism and transport carbohydrate metabolism and transport chaperone translation

668

45

gi|116495980

35 508

238

178

125  079

358  261

178

42

gi|116495240

92 48

044

185

116  040

086  033

172 106

16 14

gi|229552141 gi|199598827

49 751 76 872

205 122

091 062

101  063 114  025

168  124 183  130

Functional category

Mascot score

Coverage (%)

Accession code

chaperone chaperone

273 202

45 24

antioxidant

214

carbohydrate metabolism and transport translation

(Da);

WT = wild

type;

HS = heat-tolerant

previously obtained heat-tolerant derivative strains (du Toit et al. 2013) were determined at baseline and after 2 h of exposure to heat stress. Moreover, RT-qPCR was used to evaluate changes in the expression of the genes coding the identified proteins. This study focused on cytosolic proteins within a pI range from 4 to 7.

derivative;

BB-12 = B. animalis

HS

subsp.

lactis

BB-12;

Around 200–300 spots were detected on each of the 2D-DIGE gels. The regulation of seven spots was modified in the Bif. animalis subsp. lactis strains and another seven spots in the Lact. rhamnosus strains as a response to heat or between the original and the derivative strain (Fig. 2, Table 2 and Table 3). These spots were excised from the

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

243

Heat adaptation and response in LGG and BB-12

J. Aakko et al.

Table 3 Induction ratios of the identified proteins and the genes coding them in the heat-tolerant derivatives of Bifidobacterium animalis subsp. lactis BB-12 and Lactobacillus rhamnosus GG with regard to the wild-type strains before and after the 2 h heat treatment. For example, when the fold difference is higher than one, the induction rate of the gene or the protein is higher in the heat-tolerant derivative than in the wild-type strain Fold difference HS/WT

Strain

Spot

Protein

BB-12

BL1 BL2 BL4 BL6 BL7 LR1 LR2 LR3

Heat-shock protein 60, GroEL GroES Thioredoxin peroxidase Transketolase Elongation factor Tu Chaperonin GroEL Molecular chaperone, DnaK Glucosamine–fructose-6-phosphate aminotransferase, isomerizing L-lactate dehydrogenase Phosphocarrier protein Hpr Cell division trigger factor, peptidyl-prolyl cis-trans isomerase Elongation factor G

LGG

LR4 LR5 LR6 LR7

Mascot score

Coverage (%)

Accession number

273 202 214 127 541 419 126 570

45 24 43 27 31 37 21 52

668 178 172 106

0h

2h

2D-DIGE

RT-qPCR

2D-DIGE

RT-qPCR

gi|14719389 gi|51094292 gi|183601582 gi|289178611 gi|183602587 gi|259650517 gi|199597089 gi|258507978

16 22 32 14 09 23 19 32

29 26 18 23 29 06 07 06

16 14 19 07 04 23 11 22

32 26 30 26 18 252 135 115

45 42 16

gi|116495980 gi|116495240 gi|229552141

17 07 07

08 17 31

12 31 03

23 17 41

14

gi|199598827

09

48

05

44

HS = heat-tolerant derivative; WT = wild type; BB-12 = Bifidobacterium animalis subsp. lactis BB-12; LGG = Lactobacillus rhamnosus GG.

gels and subjected to peptide mass fingerprint analysis to identify the proteins in the spots. For spot BL3, no protein could be identified, while for spot BL5, two different proteins were detected, and thus, no conclusions can be made from the data concerning these spots. Proteins affected by heat were chaperones and proteins involved in carbohydrate metabolism and translation.

strains, while the higher levels of GroEL and thioredoxin peroxidase remained in the derivative. In addition, the derivative strain seemed to have lower levels of elongation factor Tu after the heat treatment. Anyhow, RTqPCR analysis indicated that the induction rate of all of the identified proteins was higher in the heat-tolerant derivative at both time points (Table 3).

Response of Bifidobacterium animalis subsp. lactis strains to heat

Response of Lactobacillus rhamnosus strains to heat

With regard to the response to heat, proteins identified as GroEL, GroES and thioredoxin peroxidase were upregulated in Bif. animalis subsp. lactis BB-12 WT according to the proteomic analysis (Table 2). On the contrary, RTqPCR did not show any change in the activity of the genes coding the proteins GroES and thioredoxin peroxidase, but elongation factor Tu appeared overexpressed in this strain. In the heat-tolerant derivative of Bif. animalis subsp. lactis BB-12, the heat-shock response was otherwise similar compared with the original strain, but transketolase and elongation factor Tu were downregulated according to both 2D-DIGE and RT-qPCR (Table 2). A direct comparison of the wild-type and the heat-tolerant strains (Table 3) revealed that the levels of proteins identified as GroEL, GroES and thioredoxin peroxidase were higher before the heat shock in the heat-tolerant derivative strain according to 2D-DIGE. After the heat shock, the level of GroES was comparable between the 244

According to the proteomic analysis, in Lact. rhamnosus GG WT, GroEL, L-lactate dehydrogenase and cell division trigger factor were upregulated, while glucosamine–fructose-6-phosphate aminotransferase and phosphocarrier protein Hpr were downregulated after 2 h of heat treatment when compared to baseline (Table 2). However, the RT-qPCR results suggest that the induction rate of the genes coding the proteins identified as GroEL and DnaK was lower at the 2 h time point in this strain. Likewise to the original strain, GroEL and L-lactate dehydrogenase were upregulated and glucosamine–fructose-6-phosphate aminotransferase downregulated in the heat-tolerant derivative strain of Lact. rhamnosus GG. On the contrary, the regulation of the cell division trigger factor remained stable throughout the experiment in the derivative strain, while phosphocarrier protein Hpr was upregulated and elongation factor G downregulated due to heat shock (Table 2). However, the RT-qPCR results indicated that the genes coding the chaperones GroEL and DnaK were highly expressed. Likewise, the genes

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

J. Aakko et al.

coding glucosamine–fructose-6-phosphate aminotransferase and L-lactate dehydrogenase appeared overexpressed, but to a lesser extent than the chaperones. When the protein expression profiles of the Lact. rhamnosus strains before the heat treatment were compared (Table 3), higher levels of GroEL, DnaK, glucosamine–fructose-6-phosphate aminotransferase and Llactate dehydrogenase were detected in the heat-tolerant derivative strain. However, these results were not confirmed by the RT-qPCR results, which indicated a slight underexpression of the genes coding GroEL, DnaK, glucosamine–fructose-6-phosphate aminotransferase and Llactate dehydrogenase, while phosphocarrier protein Hpr, cell division trigger factor and elongation factor G were overexpressed in the derivative strain compared with the original strain (Table 3). After the heat treatment, proteomic results indicated that the derivative strain still had higher levels of GroEL, glucosamine–fructose-6-phosphate aminotransferase and lower levels of trigger factor and elongation factor G than the wild-type strain (Table 3). Furthermore, the protein Hpr was induced at a higher level in the derivative strain, while both strains had comparable levels of DnaK and L-lactate dehydrogenase (Table 3). However, these results were not confirmed by the RT-qPCR, which indicated a higher activity of all the genes measured at 2 h in the derivative strain. Discussion Heat-shock response in Bifidobacterium animalis subsp. lactis strains According to proteomic studies, chaperones GroEL and GroES, which are known to work together as a complex (van de Guchte et al. 2002), were upregulated by heat in both Bif. animalis subsp. lactis BB-12 strains. However, the expression level of the gene coding GroES did not change according to RT-qPCR. Chaperones are proteins that facilitate the folding of newly synthesized proteins and refold denatured ones and thus protect the host cells from stresses. In lactobacilli and bifidobacteria, chaperones have been reported as one of the major stress response systems against heat stress (Sugimoto et al. 2008). GroEL and GroES are known as one of the classical heat-shock proteins, but they are also known to be a part of the general stress response in bifidobacteria and lactobacilli (Kilstrup et al. 1997; Prasad et al. 2003; Sanchez et al. 2005; Ventura et al. 2005), and many stress response mechanisms have been described to be dependent on the GroEL and GroES activity in a stress regulatory network model for bifidobacteria (Zomer et al. 2009). Although the induction rate of the chaperones was similar between the Bif. animalis subsp. lactis strains, the

Heat adaptation and response in LGG and BB-12

heat-tolerant derivative of BB-12 harboured higher levels of both GroEL and GroES before and after the treatment. This might be one of the factors responsible for the enhanced heat tolerance in the derivative strain. Transketolase, a protein involved in carbohydrate metabolism, appeared downregulated in the heat-tolerant derivative strain. This protein is an enzyme responsible for the formation of xylulose 5-phosphate that is one intermediate in the so-called bifidobacterial shunt, which is a unique pathway that bifidobacteria use to utilize hexoses (Sanchez et al. 2005). One translational protein, elongation factor Tu (EFTu), was downregulated in the heat-tolerant derivative of Bif. animalis subsp. lactis BB-12 according to both 2DDIGE and RT-qPCR. In the wild-type strain, the regulation of this protein remained stable, but the RT-qPCR results indicated that the gene coding this protein was slightly upregulated as a response to heat. EF-Tu has been reported to be induced by acid in Propionibacterium freudenreichii (Leverrier et al. 2004), by bile in Bif. longum (Sanchez et al. 2005) and by acid stress in Streptococcus mutans (Wu et al. 2011). Elongation factors are involved in the delivery of aminoacyl-tRNA to the ribosome (Stark et al. 2002), and they have as well been described as chaperones on unfolded and denatured proteins in E. coli (Caldas et al. 2000). Hence, the overexpression of the gene in the wild-type strain during heat shock would be logical, but the reason why the regulation of this particular protein does not seem to change remains unknown. Interestingly, it has been noticed that EF-Tu might appear in different isoforms after environmental stress, most likely due to post-translational modifications such as phosphorylation (Wilkins et al. 2002; Leverrier et al. 2004). If EF-Tu appears in different isoforms having different functions, it can also appear as more than one spot on the 2D-PAGE/DIGE gels, and thus, we might be looking at only one of the isoforms in our proteomic studies. Moreover, elongation factors may also appear extracellular or in cell walls (Izquierdo et al. 2009), but in the present study, only the cytoplasmic proteins were analysed, which could lead to an underestimation of the regulation by the used methods. An antioxidant, thioredoxin peroxidase, was identified as upregulated in the Bif. animalis subsp. lactis BB-12 strains according to the proteomic studies, and higher levels of this protein were detected in the heat-tolerant derivative than in the wild-type strain before and after the heat treatment. This protein has been described as one of the main cellular enzymes for the detoxification of H2O2 and as a specific inducer of thioredoxin and thioredoxin reductase gene expression in response to oxidative stress in Saccharomyces cerevisae (Ross et al. 2000). In addition, thioredoxin peroxidase has been suggested to be

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

245

Heat adaptation and response in LGG and BB-12

J. Aakko et al.

involved in stress response in Bif. animalis subsp. lactis (Sanchez et al. 2007) and B. subtilis (Scharf et al. 1998). Thus, we believe that this protein may help the heat-tolerant derivative of Bif. animalis subsp. lactis BB-12 to survive better under heat shock. Heat-shock response in Lactobacillus rhamnosus strains Likewise to B. lactis strains, the regulation of two proteins identified as chaperones, GroEL and DnaK, was affected by the heat treatment. GroEL was upregulated in both strains, but the high induction rate of DnaK was only detected by RT-qPCR in the heat-tolerant derivative. On the other hand, it is possible that our proteomic assay underestimates the amount of DnaK because the surface proteins were excluded from the analysis. For example, DnaK has been reported to be induced by bile in Bif. animalis subsp. lactis when the cell wall proteome was examined (Candela et al. 2010). One mechanism responsible for the overexpression of DnaK has been described in a heat-tolerant Bifidobacterium longum strain obtained in a similar fashion to our strains (Berger et al. 2010). In this strain, point mutations were found in the hspR gene, which codes a negative regulator of DnaK and clpB. The ability to induce chaperones (GroEL and DnaK) at higher concentrations in the heat-tolerant derivative Lact. rhamnosus GG HS strain might contribute to the increased heat resistance. A protein involved in glutamate metabolism (glucosamine–fructose-6-phosphate aminotransferase, isomerizing) was downregulated in the Lact. rhamnosus strains during heat stress. However, the gene coding this protein was overexpressed in the heat-tolerant derivative according to the RT-qPCR results. This enzyme has been reported to help the membrane recovery after damage caused by environmental stresses (Milewski 2002; Wu et al. 2011). Two proteins involved in carbohydrate metabolism were affected by the heat treatment in Lact. rhamnosus srains: phosphocarrier protein Hpr and L-lactate dehydrogenase. The phosphocarrier protein Hpr, a component of the phosphoenolpyruvate-dependent sugar phosphotransferase system (van Nuland et al. 1995), was downregulated in the original strain, but upregulated in the heattolerant derivative according to the proteomic studies. However, the RT-qPCR results did not show any change of regulation. Hpr may also exist in a phosphorylated form (Schmidl et al. 2010), so it is possible that the observed up- or downregulation in the proteomic studies is caused by a difference in the phosphorylation status of the protein. The Hpr protein is known to negatively regulate the protease production and has been described to help B. subtilis survive from oxidative stress (Dowds and Hoch 1991). Earlier, Hpr has been revealed upregulated 246

during heat and osmotic stress in Lact. rhamnosus HN001 (Prasad et al. 2003). Moreover, the phosphotransferase system might play a role in cold-shock response of L. casei, although the mechanism implied remains to be elucidated (Monedero et al. 2007). In addition to Hpr, another protein involved in carbohydrate metabolism, L-lactate dehydrogenase, was upregulated after 2 h of heat treatment in the Lact. rhamnosus strains according to the proteomic results. However, the gene expression was induced at a higher extent in the heat-tolerant derivative. The lactate dehydrogenase has been reported to be induced by bile exposure in Bif. longum (Sanchez et al. 2005), various stresses in Enterococcus faecalis (Giard et al. 2001), by heat in Lactobacillus gasseri (Suokko et al. 2008), and help Neisseria gonorrhoeae to survive under oxidative stress (Fu et al. 1989). The cytoplasmic L-lactate dehydrogenase is NAD-dependent, and it catalyses the conversion of pyruvate to lactate, which is a major end product of carbohydrate fermentation in lactic acid bacteria and bifidobacteria (Fu et al. 1989; Sanchez et al. 2005). According to proteomic studies, the regulation of two translational proteins was modified in the Lact. rhamnosus strains following the heat treatment. A protein identified as elongation factor G (EF-G) was downregulated in the heattolerant derivative following the heat treatment. Earlier, this protein has been reported to be induced by acid stress in Strep. mutans (Wu et al. 2011). The second translational protein, cell division trigger factor, was upregulated in the original strain after 2 h of heat treatment in the proteomic studies, although the expression of the gene encoding the protein barely changed according to the RT-qPCR results. This protein has been reported to work together with enzymes involved in protein folding (Schonbrunner and Schmid 1992; Gothel and Marahiel 1999). In our study, some discrepancies were observed between the protein and gene expression analyses. Similar discrepancies have been reported by other authors as well when the stress response in lactic acid bacteria has been examined using the aforementioned techniques (Wu et al. 2011). These may be related to the different sensitivities of the techniques, but it must also be kept in mind that RT-qPCR quantifies the mRNA, while proteomic techniques quantify the protein in the cell, and due to post-translational regulation and modifications, the two techniques may occasionally give contradictory results (Hegde et al. 2003). In addition, we only analysed the cytoplasmic proteins, which may affect the outcome as some of the proteins analysed have been reported to appear in the cell wall as well. Taken together, the acquisition of heat tolerance may be associated with at least two different mechanisms, a slightly increased constitutive level of chaperones in B. lactis and

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

J. Aakko et al.

an increased ability to induce their production in Lact. rhamnosus. Identifying conditions in which the identified chaperones are produced at higher levels might increase the heat tolerance of the study strains even further and allow their inclusion in new probiotic food categories. Similar changes have been observed before and during various stresses, so the acquisition of heat tolerance might affect the tolerance to other stress factors as well. However, it may change for better or for worse as described in Bif. longum for instance (Mozzetti et al. 2013), and there may be some other unexpected and maybe unwanted genetic or other type of changes that may affect the probiotic potential of the strain. Thus, the verification of the presence of the original probiotic properties in the heattolerant derivative strains should be confirmed to enable their use in functional foods and in products aimed at health benefits for humans. Moreover, the impact of the processing conditions, production processes and carrier vehicle, either food, food component or pharmaceutical component, on the probiotic properties should also be routinely evaluated (Grzeskowiak et al. 2011). Acknowledgements This study was funded by TEKES (the Finnish Funding Agency for Technology and Innovation) grant number 398/31/2009. Conflict of interest No conflict of interest declared. References Barreiro, C., Gonzalez-Lavado, E., Brand, S., Tauch, A. and Martin, J.F. (2005) Heat shock proteome analysis of wildtype Corynebacterium glutamicum ATCC 13032 and a spontaneous mutant lacking GroEL1, a dispensable chaperone. J Bacteriol 187, 884–889. Berger, B., Moine, D., Mansourian, R. and Arigoni, F. (2010) HspR mutations are naturally selected in Bifidobacterium longum when successive heat shock treatments are applied. J Bacteriol 192, 256–263. Caldas, T., Laalami, S. and Richarme, G. (2000) Chaperone properties of bacterial elongation factor EF-G and initiation factor IF2. J Biol Chem 275, 855–860. Candela, M., Centanni, M., Fiori, J., Biagi, E., Turroni, S., Orrico, C., Bergmann, S., Hammerschmidt, S. et al. (2010) DnaK from Bifidobacterium animalis subsp. lactis is a surface-exposed human plasminogen receptor upregulated in response to bile salts. Microbiology 156, 1609–1618. De Angelis, M. and Gobbetti, M. (2004) Environmental stress responses in Lactobacillus: a review. Proteomics 4, 106–122.

Heat adaptation and response in LGG and BB-12

De Man, J.C., Rogosa, M. and Sharpe, M.E. (1960) A medium for the cultivation of lactobacilli. J Appl Bacteriol 23, 130–135. Dominguez, F., Garrido-Gomez, T., Lopez, J.A., Camafeita, E., Quinonero, A., Pellicer, A. and Simon, C. (2009) Proteomic analysis of the human receptive versus non-receptive endometrium using differential in-gel electrophoresis and MALDI-MS unveils stathmin 1 and annexin A2 as differentially regulated. Hum Reprod 24, 2607–2617. Dowds, B.C. and Hoch, J.A. (1991) Regulation of the oxidative stress response by the hpr gene in Bacillus subtilis. J Gen Microbiol 137, 1121–1125. FAO/WHO (2002) Report of a Joint FAO/WHO Working Group on Drafting Guidelines for the Evaluation of Probiotics in Food. London, Canada: Joint FAO/WHO Working Group on Drafting Guidelines for the Evaluation of Probiotics in Food. Fu, H.S., Hassett, D.J. and Cohen, M.S. (1989) Oxidant stress in Neisseria gonorrhoeae: adaptation and effects on L-(+)-lactate dehydrogenase activity. Infect Immun 57, 2173–2178. Gao, H., Wang, Y., Liu, X., Yan, T., Wu, L., Alm, E., Arkin, A., Thompson, D.K. et al. (2004) Global transcriptome analysis of the heat shock response of Shewanella oneidensis. J Bacteriol 186, 7796–7803. Giard, J.C., Laplace, J.M., Rince, A., Pichereau, V., Benachour, A., Leboeuf, C., Flahaut, S., Auffray, Y. et al. (2001) The stress proteome of Enterococcus faecalis. Electrophoresis 22, 2947–2954. Gothel, S.F. and Marahiel, M.A. (1999) Peptidyl-prolyl cistrans isomerases, a superfamily of ubiquitous folding catalysts. Cell Mol Life Sci 55, 423–436. Grzeskowiak, L., Isolauri, E., Salminen, S. and Gueimonde, M. (2011) Manufacturing process influences properties of probiotic bacteria. Br J Nutr 105, 887–894. van de Guchte, M., Serror, P., Chervaux, C., Smokvina, T., Ehrlich, S.D. and Maguin, E. (2002) Stress responses in lactic acid bacteria. Antonie Van Leeuwenhoek 82, 187–216. Hegde, P.S., White, I.R. and Debouck, C. (2003) Interplay of transcriptomics and proteomics. Curr Opin Biotechnol 14, 647–651. Helmann, J.D., Wu, M.F., Kobel, P.A., Gamo, F.J., Wilson, M., Morshedi, M.M., Navre, M. and Paddon, C. (2001) Global transcriptional response of Bacillus subtilis to heat shock. J Bacteriol 183, 7318–7328. Izquierdo, E., Horvatovich, P., Marchioni, E., Aoude-Werner, D., Sanz, Y. and Ennahar, S. (2009) 2-DE and MS analysis of key proteins in the adhesion of Lactobacillus plantarum, a first step toward early selection of probiotics based on bacterial biomarkers. Electrophoresis 30, 949–956. Kilstrup, M., Jacobsen, S., Hammer, K. and Vogensen, F.K. (1997) Induction of heat shock proteins DnaK, GroEL, and GroES by salt stress in Lactococcus lactis. Appl Environ Microbiol 63, 1826–1837. Leverrier, P., Vissers, J.P., Rouault, A., Boyaval, P. and Jan, G. (2004) Mass spectrometry proteomic analysis of stress adaptation reveals both common and distinct response

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

247

Heat adaptation and response in LGG and BB-12

J. Aakko et al.

pathways in Propionibacterium freudenreichii. Arch Microbiol 181, 215–230. Livak, K.J. and Schmittgen, T.D. (2001) Analysis of relative gene expression data using real-time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25, 402–408. Milewski, S. (2002) Glucosamine-6-phosphate synthase–the multi-facets enzyme. Biochim Biophys Acta 1597, 173–192. Monedero, V., Maze, A., Boel, G., Zuniga, M., Beaufils, S., Hartke, A. and Deutscher, J. (2007) The phosphotransferase system of Lactobacillus casei: regulation of carbon metabolism and connection to cold shock response. J Mol Microbiol Biotechnol 12, 20–32. Mozzetti, V., Grattepanche, F., Berger, B., Rezzonico, E., Arigoni, F. and Lacroix, C. (2013) Fast screening of Bifidobacterium longum sublethal stress conditions in a novel two-stage continuous culture strategy. Benef Microbes 4, 167–178. van Nuland, N.A., Boelens, R., Scheek, R.M. and Robillard, G.T. (1995) High-resolution structure of the phosphorylated form of the histidine-containing phosphocarrier protein HPr from Escherichia coli determined by restrained molecular dynamics from NMRNOE data. J Mol Biol 246, 180–193. Prasad, J., McJarrow, P. and Gopal, P. (2003) Heat and osmotic stress responses of probiotic Lactobacillus rhamnosus HN001 (DR20) in relation to viability after drying. Appl Environ Microbiol 69, 917–925. Rezzonico, E., Lariani, S., Barretto, C., Cuanoud, G., Giliberti, G., Delley, M., Arigoni, F. and Pessi, G. (2007) Global transcriptome analysis of the heat shock response of Bifidobacterium longum. FEMS Microbiol Lett 271, 136–145. Richmond, C.S., Glasner, J.D., Mau, R., Jin, H. and Blattner, F.R. (1999) Genome-wide expression profiling in Escherichia coli K-12. Nucleic Acids Res 27, 3821–3835. Ross, S.J., Findlay, V.J., Malakasi, P. and Morgan, B.A. (2000) Thioredoxin peroxidase is required for the transcriptional response to oxidative stress in budding yeast. Mol Biol Cell 11, 2631–2642. Sanchez, B., Champomier-Verges, M.C., Anglade, P., Baraige, F., de Los Reyes-Gavilan, C.G., Margolles, A. and Zagorec, M. (2005) Proteomic analysis of global changes in protein expression during bile salt exposure of Bifidobacterium longum NCIMB 8809. J Bacteriol 187, 5799–5808. Sanchez, B., Champomier-Verges, M.C., Stuer-Lauridsen, B., Ruas-Madiedo, P., Anglade, P., Baraige, F., de los ReyesGavilan, C.G., Johansen, E. et al. (2007) Adaptation and response of Bifidobacterium animalis subsp. lactis to bile: a proteomic and physiological approach. Appl Environ Microbiol 73, 6757–6767. Scharf, C., Riethdorf, S., Ernst, H., Engelmann, S., Volker, U. and Hecker, M. (1998) Thioredoxin is an essential protein induced by multiple stresses in Bacillus subtilis. J Bacteriol 180, 1869–1877. Schmidl, S.R., Gronau, K., Pietack, N., Hecker, M., Becher, D. and Stulke, J. (2010) The phosphoproteome of the minimal

248

bacterium Mycoplasma pneumoniae: analysis of the complete known Ser/Thr kinome suggests the existence of novel kinases. Mol Cell Proteomics 9, 1228–1242. Schonbrunner, E.R. and Schmid, F.X. (1992) Peptidyl-prolyl cis-trans isomerase improves the efficiency of protein disulfide isomerase as a catalyst of protein folding. Proc Natl Acad Sci USA 89, 4510–4513. Stark, H., Rodnina, M.V., Wieden, H.J., Zemlin, F., Wintermeyer, W. and van Heel, M. (2002) Ribosome interactions of aminoacyl-tRNA and elongation factor Tu in the codon-recognition complex. Nat Struct Biol 9, 849– 854. Stewart, G.R., Wernisch, L., Stabler, R., Mangan, J.A., Hinds, J., Laing, K.G., Young, D.B. and Butcher, P.D. (2002) Dissection of the heat-shock response in Mycobacterium tuberculosis using mutants and microarrays. Microbiology 148, 3129–3138. Sugimoto, S., Abdullah-Al-Mahin, and Sonomoto, K. (2008) Molecular chaperones in lactic acid bacteria: physiological consequences and biochemical properties. J Biosci Bioeng 106, 324–336. Suokko, A., Poutanen, M., Savijoki, K., Kalkkinen, N. and Varmanen, P. (2008) ClpL is essential for induction of thermotolerance and is potentially part of the HrcA regulon in Lactobacillus gasseri. Proteomics 8, 1029–1041. du Toit, E., Vesterlund, S., Gueimonde, M. and Salminen, S. (2013) Assessment of the effect of stress-tolerance acquisition on some basic characteristics of specific probiotics. Int J Food Microbiol 165, 51–56. Tuomola, E., Crittenden, R., Playne, M., Isolauri, E. and Salminen, S. (2001) Quality assurance criteria for probiotic bacteria. Am J Clin Nutr 73, 393S–398S. Ventura, M., Zink, R., Fitzgerald, G.F. and van Sinderen, D. (2005) Gene structure and transcriptional organization of the dnaK operon of Bifidobacterium breve UCC 2003 and application of the operon in bifidobacterial tracing. Appl Environ Microbiol 71, 487–500. Weiner, J. 3rd, Zimmerman, C.U., Gohlmann, H.W. and Herrmann, R. (2003) Transcription profiles of the bacterium Mycoplasma pneumoniae grown at different temperatures. Nucleic Acids Res 31, 6306–6320. Wessel, D. and Flugge, U.I. (1984) A method for the quantitative recovery of protein in dilute solution in the presence of detergents and lipids. Anal Biochem 138, 141–143. Wilkins, J.C., Homer, K.A. and Beighton, D. (2002) Analysis of Streptococcus mutans proteins modulated by culture under acidic conditions. Appl Environ Microbiol 68, 2382–2390. Wu, R., Zhang, W., Sun, T., Wu, J., Yue, X., Meng, H. and Zhang, H. (2011) Proteomic analysis of responses of a new probiotic bacterium Lactobacillus casei Zhang to low acid stress. Int J Food Microbiol 147, 181–187. Zomer, A., Fernandez, M., Kearney, B., Fitzgerald, G.F., Ventura, M. and van Sinderen, D. (2009) An interactive regulatory network controls stress response in Bifidobacterium breve UCC2003. J Bacteriol 191, 7039–7049.

Journal of Applied Microbiology 117, 239--248 © 2014 The Society for Applied Microbiology

Assessment of stress tolerance acquisition in the heat-tolerant derivative strains of Bifidobacterium animalis subsp. lactis BB-12 and Lactobacillus rhamnosus GG.

The purpose of this study was to investigate the heat-shock response at molecular level in Lactobacillus rhamnosus GG, Bifidobacterium animalis subsp...
219KB Sizes 0 Downloads 3 Views