8
Multigenic expression analysis as an approach to understanding the behaviour of Oenococcus oeni in wine-like conditions Nair Olguín, Albert Bordons, Cristina Reguant Departament de Bioquímica i Biotecnologia, Facultat d'Enologia, Universitat Rovira i Virgili, C/Marcel·lí Domingo s/n 43007, Tarragona, Catalonia, Spain abstract article info Article history: Received 9 June 2010 Received in revised form 26 August 2010 Accepted 31 August 2010 Keywords: Oenococcus oeni Malolactic fermentation Wine RT qPCR Gene expression The correct performance of wine malolactic fermentation (MLF) depends on the metabolic characteristics of the Oenococcus oeni strain/s responsible for this process. This study characterizes four O. oeni strains, which behave differently in terms of malolactic performance, the citric acid use related to acetic acid production, and stress adaptation. Metabolic evolution and its associated enzymatic activities were studied and the transcriptional response of the genes related to MLF, citrate metabolism and stress response was compared among strains. A higher initial expression of both the malolactic enzyme and the encoding gene mleA may be related to faster MLF. The initial transcriptional levels of citrate lyase (citE) proved indicative of early citrate consumption. Moreover, the strains that performed best in wine-like conditions presented a much higher relative expression of several stress responsive genes, particularly hsp18, clpP, ctsR and rmlB. Finally, an inter- strain comparison of the transcriptional levels of selected genes at different times during MLF proved a useful tool in characterizing strains based on their metabolic behaviour. © 2010 Elsevier B.V. All rights reserved. 1. Introduction Oenococcus oeni is the best adapted wine-associated lactic acid bacteria and is almost exclusively used for the induction of malolactic fermentation (MLF) in red, white and sparkling base wines (Wibowo et al., 1985). MLF is the bacterial-driven decarboxylation of L-malic acid to L-lactic acid and carbon dioxide, resulting in deacidication, avour modications and the microbial stability of wine (Bartowsky, 2005). In wine, O. oeni comes up against multiple sources of stress such as ethanol, sulphur dioxide concentration, nutrient depletion, temperature and low pH, which inuence population development (Vaillant et al., 1995; Versari et al., 1999). In order to survive in such environmental conditions, O. oeni uses several mechanisms that have been associated with a possible stress response and adaptation. The metabolism of malate and citrate involves proton consumption and the generation of both membrane potential and pH gradient (Augagneur et al., 2007; Cox and Henick- Kling, 1989; Loubiere et al., 1992; Salema et al., 1996), thus allowing ATP synthesis and the deacidication of intracellular and extracellular media. Moreover, citrate utilization leads to the production of avouring compounds, such as diacetyl, acetoin, butanediol and acetate, thus contributing to the organoleptic characteristics of wine (Ramos and Santos, 1996). Until recently, little was known about malate and citrate metabolism at the molecular level. Today, and thanks to the genome sequence of O. oeni PSU-1 (Genbank NC_008528) it is possible to investigate the genes involved in this pathway and their level of transcription under different growth conditions (Augagneur et al., 2007; Beltramo et al., 2006; Olguín et al., 2009). Another mechanism involved in O. oeni adaptation to low pH and the presence of ethanol is adjusting membrane uidity by modifying its composition (Da Silveira et al., 2003; Grandvalet et al., 2008; Teixeira et al., 2002). For instance, O. oeni cells respond to culture in the presence of ethanol by increasing their cyclopropane fatty acid (CFA) content, which is thought to counteract the effect of ethanol on membrane uidity (Teixeira et al., 2002). In addition, the increase in CFA content observed in acid- and ethanol-grown cells has also been related to a higher level of cfa transcripts, suggesting the transcrip- tional regulation of this gene under different stress conditions (Grandvalet et al., 2008). The synthesis of heat-shock proteins, chaperones and proteases by O. oeni under stress conditions has also been studied (Beltramo et al., 2004; Guzzo et al., 1997, 2000; Jobin et al., 1997). Recent studies have conrmed the transcriptional regulation of the small heat-shock protein Lo18 at the molecular level (Coucheney et al., 2005). Additionally, O. oeni cells exhibiting an increased synthesis of Lo18 have shown a greater ability to survive in wine and to perform MLF. Furthermore, a multigenic analysis has recently been conducted which aimed to quantify the transcriptional level of 13 genes that could be involved in O. oeni adaptation to wine conditions (Beltramo et al., 2006). Using one strain of O. oeni these authors assessed the overall response to stress during MLF. However, other O. oeni strains International Journal of Food Microbiology 144 (2010) 8895 Corresponding author. Tel.: + 34 977 558280; fax: + 34 977 558232. E-mail address: [email protected] (C. Reguant). 0168-1605/$ see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.ijfoodmicro.2010.08.032 Contents lists available at ScienceDirect International Journal of Food Microbiology journal homepage: www.elsevier.com/locate/ijfoodmicro

Olguin Et Al. 2010

Embed Size (px)

DESCRIPTION

Multigenic expression analysis as an approach to understanding the behaviour of Oenococcus oeni in wine-like conditions

Citation preview

Page 1: Olguin Et Al. 2010

International Journal of Food Microbiology 144 (2010) 88–95

Contents lists available at ScienceDirect

International Journal of Food Microbiology

j ourna l homepage: www.e lsev ie r.com/ locate / i j foodmicro

Multigenic expression analysis as an approach to understanding the behaviour ofOenococcus oeni in wine-like conditions

Nair Olguín, Albert Bordons, Cristina Reguant ⁎Departament de Bioquímica i Biotecnologia, Facultat d'Enologia, Universitat Rovira i Virgili, C/Marcel·lí Domingo s/n 43007, Tarragona, Catalonia, Spain

⁎ Corresponding author. Tel.: +34 977 558280; fax:E-mail address: [email protected] (C. Reguan

0168-1605/$ – see front matter © 2010 Elsevier B.V. Adoi:10.1016/j.ijfoodmicro.2010.08.032

a b s t r a c t

a r t i c l e i n f o

Article history:Received 9 June 2010Received in revised form 26 August 2010Accepted 31 August 2010

Keywords:Oenococcus oeniMalolactic fermentationWineRT qPCRGene expression

The correct performance of wine malolactic fermentation (MLF) depends on the metabolic characteristics ofthe Oenococcus oeni strain/s responsible for this process. This study characterizes four O. oeni strains, whichbehave differently in terms of malolactic performance, the citric acid use related to acetic acid production, andstress adaptation. Metabolic evolution and its associated enzymatic activities were studied and thetranscriptional response of the genes related to MLF, citrate metabolism and stress response was comparedamong strains. A higher initial expression of both the malolactic enzyme and the encoding gene mleA may berelated to faster MLF. The initial transcriptional levels of citrate lyase (citE) proved indicative of early citrateconsumption. Moreover, the strains that performed best in wine-like conditions presented a much higherrelative expression of several stress responsive genes, particularly hsp18, clpP, ctsR and rmlB. Finally, an inter-strain comparison of the transcriptional levels of selected genes at different times during MLF proved a usefultool in characterizing strains based on their metabolic behaviour.

+34 977 558232.t).

ll rights reserved.

© 2010 Elsevier B.V. All rights reserved.

1. Introduction

Oenococcus oeni is the best adapted wine-associated lactic acidbacteria and is almost exclusively used for the induction of malolacticfermentation (MLF) in red, white and sparkling base wines (Wibowoet al., 1985). MLF is the bacterial-driven decarboxylation of L-malicacid to L-lactic acid and carbon dioxide, resulting in deacidification,flavour modifications and the microbial stability of wine (Bartowsky,2005). In wine, O. oeni comes up against multiple sources of stresssuch as ethanol, sulphur dioxide concentration, nutrient depletion,temperature and low pH, which influence population development(Vaillant et al., 1995; Versari et al., 1999).

In order to survive in such environmental conditions, O. oeni usesseveral mechanisms that have been associated with a possible stressresponse and adaptation. The metabolism of malate and citrateinvolves proton consumption and the generation of both membranepotential and pH gradient (Augagneur et al., 2007; Cox and Henick-Kling, 1989; Loubiere et al., 1992; Salema et al., 1996), thus allowingATP synthesis and the deacidification of intracellular and extracellularmedia. Moreover, citrate utilization leads to the production offlavouring compounds, such as diacetyl, acetoin, butanediol andacetate, thus contributing to the organoleptic characteristics of wine(Ramos and Santos, 1996). Until recently, little was known aboutmalate and citrate metabolism at the molecular level. Today, and

thanks to the genome sequence of O. oeni PSU-1 (GenbankNC_008528) it is possible to investigate the genes involved in thispathway and their level of transcription under different growthconditions (Augagneur et al., 2007; Beltramo et al., 2006; Olguín et al.,2009).

Another mechanism involved in O. oeni adaptation to low pH andthe presence of ethanol is adjusting membrane fluidity by modifyingits composition (Da Silveira et al., 2003; Grandvalet et al., 2008;Teixeira et al., 2002). For instance, O. oeni cells respond to culture inthe presence of ethanol by increasing their cyclopropane fatty acid(CFA) content, which is thought to counteract the effect of ethanol onmembrane fluidity (Teixeira et al., 2002). In addition, the increase inCFA content observed in acid- and ethanol-grown cells has also beenrelated to a higher level of cfa transcripts, suggesting the transcrip-tional regulation of this gene under different stress conditions(Grandvalet et al., 2008).

The synthesis of heat-shock proteins, chaperones and proteases byO. oeni under stress conditions has also been studied (Beltramo et al.,2004; Guzzo et al., 1997, 2000; Jobin et al., 1997). Recent studies haveconfirmed the transcriptional regulation of the small heat-shockprotein Lo18 at the molecular level (Coucheney et al., 2005).Additionally, O. oeni cells exhibiting an increased synthesis of Lo18have shown a greater ability to survive in wine and to perform MLF.Furthermore, a multigenic analysis has recently been conductedwhich aimed to quantify the transcriptional level of 13 genes thatcould be involved in O. oeni adaptation to wine conditions (Beltramoet al., 2006). Using one strain of O. oeni these authors assessed theoverall response to stress during MLF. However, other O. oeni strains

Page 2: Olguin Et Al. 2010

89N. Olguín et al. / International Journal of Food Microbiology 144 (2010) 88–95

as well as other genes need to be studied in order to better understandthe level of gene induction and cell adaptation mechanisms. Inkeeping with this, evaluating the metabolic pathways at thetranscriptional level could also contribute to understanding bothcell adaptation and secondary metabolite production.

The aim of this work was to evaluate the relationship between thetranscriptional and functional behaviours of several stress responsiveandmetabolic related genes under wine-like conditions in different O.oeni strains. The expression of twelve genes was quantified in fourstrains of O. oeni by means of reverse transcription real-timequantitative polymerase chain reaction (RT qPCR). The productionof metabolites from citrate utilization as well as the activity of relatedenzymes was monitored to evaluate its correlation with theexpression of the corresponding genes. MLF was also evaluated bymeans of both enzyme activity and gene expression.

2. Materials and methods

2.1. Oenococcus oeni strains

The strains used in this study were selected from a previousscreening based on their metabolic behaviour. Four strains werechosen due to their differences in growth rate, malate and citrateconsumption time and acetate production: PSU-1 (ATCC BAA-331),CECT 217T (type strain), Microenos (MO, Laffort, S.A.) and Z42 (owncollection).

2.2. Growth conditions

O. oeni strains were grown in MRS broth medium supplementedwith D,L-malic acid (8 g/L) and fructose (5 g/L) at pH 5.0. Cells werecollected at the end of the exponential phase (OD600nm=1) andinoculated after centrifugation to a final concentration of 107 CFU/mlin 500 ml flasks containing cFT80 medium (Olguín et al., 2009) at pH4.0 modified by the addition of 1 g/L of citric acid. Following the sameprocedure, when cultures again reached the end of the exponentialphase, they were inoculated in 2 l flasks containing cFT80 medium atpH 3.5 with 12% (vol/vol) ethanol and a reduced amount of sugar(glucose 1 g/L and fructose 1 g/L). The incubation conditions werealways at 28 °C in a CO2 incubator. All assays were performed induplicate and growth was monitored by measuring absorbance andcounting plates in MRS medium (De Man et al., 1960) supplementedwith L-malic acid (4 g/L) and fructose (5 g/L) at pH 5.0.

2.3. RNA extraction

Cells were harvested by centrifugation, frozen in liquid nitrogenand kept at −80 °C until RNA extraction. Total RNA extractions wereperformed as described Chomczynski and Sacchi (1987) and thenpurified using a Roche RNeasy kit according to the manufacturer's

Table 1Primers used in this work.

Target gene Description Forward primer (5′→3′)

mleA Malolactic enzyme CCGACAATTGCTGATACAATTGAAcitE Citrate lyase β subunit CCGCACGATGATGTTTGTTCCackA Acetate kinase GCTGATGCGCTTGTTTTCACGalsD α Acetolactate decarboxylase GCCGCAATTAGAGTACACGddl D:alanine–D:alanine ligase CCTAAGCGAGCCTAATACACrmlB dtdP-glucose-4,6-dehydratase TATCCGCAATGCGCAATTGGcfa Cyclopropane fatty acid synthase TGGTATTACATTGAGCGAGGAGctsR Master regulator of stress response GGGCCATGGCAGAAGCTAATATTTCAGhsp18 Stress protein Lo18 CGGTATCAGGAGTTTTGAGTTCclpX Clp ATPase protein GAAGCGTGTTAACGAGTCCclpP ClpP protease CGGTACCAAAGGCAAGCGTTTTATtrxA Thioredoxin GCCACTTGGTGTACCCCTTGTldhD D-lactate dehydrogenase GCCGCAGTAAAGAACTTGATG

instructions (Mannheim). RNA concentrations were calculated bymeasuring absorbance at 260 nm using a Thermospectronic Genesys10 UV Spectrophotometer (Thermo Fisher Scientific).

2.4. Nucleotide sequences

Nucleotide sequences for primer design in this work wereobtained from the National Center for Biotechnology Information(NCBI). The sequence references of the genes from O. oeni PSU-1(NC_008528) are the following: ddl OEOE_0672 (YP_810247), rmlBOEOE_1447 (YP_810973) and clpX (YP_811238). The other primersused had been designed in previous works, as indicated in Table 1.

2.5. Reverse transcription and real-time qPCR

cDNA was synthesized from RNA (15 ng/μL) using TaqManReverse Transcription Reagents (Applied Biosystems) as recom-mended. Three pairs of primers were designed (Table 1) to be about18–22 bases long, to contain over 50% G/C and to have a meltingtemperature (Tm) above 60 °C. The length of the PCR products rangedfrom 90 to 156 bp. Ten pairs of primers were taken from previousworks (Beltramo et al., 2006; Desroche et al., 2005; Olguín et al., 2008,2009). The O. oeni ldhD gene, coding for lactic acid dehydrogenase,was used as an internal control, as described Desroche et al. (2005).

Real-time PCR was performed in 25 μl final volume containing7.5 ng of total cDNA, 1.5 μl of each primer at 5 μM, 4.5 μl of RNAsa freewater and 12.5 μl of SYBR Green Master Mix (Applied Biosystems).Amplifications were carried out using a Real Time PCR System 7300(Applied Biosystems) with an initial step at 95 °C for 10 min followedby 40 cycles of 95 °C for 15 s, 60 °C for 1 min and 72 °C for 30 s. Anadditional step starting between 90 and 60 °C was performed toestablish a melting curve, and was used to verify the specificity of thereal-time PCR reaction for each primer pair.

The threshold value used in this study was automaticallydetermined by the instrument. Results were analyzed using thecomparative critical threshold (ΔΔCT) method in which the amountof target RNA was adjusted to a reference (internal target RNA) aspreviously described (Livak and Schmittgen, 2001). The results showthe analysis of cDNA samples of duplicated assays, started fromindependent cultures. For each cDNA sample, each gene was analyzedin triplicate.

2.6. Inter-strain comparison of relative gene expression data

In order to compare the transcriptional response of the studiedgenes among the assayed strains and evaluate its relationship withthe different malolactic and metabolic behaviours, the relativeexpression level (RE) was represented in some cases as an inter-strain comparison. This means that the RE was calculated using thestrain showing the lowest relative expression for the studied gene at

Reverse primer (5′→3′) Amplicon length (bp) Reference

GGCATCAGAAACGACCAGCAG 156 Beltramo et al. (2006)GCTCAAAGAAACGGCATCTTCC 108 Olguín et al. (2009)AATGCCCAAGAAAGCCAGACG 90 Olguín et al. (2009)CGCGACCTTTGAAAATGGC 93 Olguín et al. (2009)CCTGTTGGCTTCGATTCATTCC 92 This workGAACGGTCAACATGCGATTCAG 93 This workCGTCTTTGAGATCACGATAATCC 113 Beltramo et al. (2006)CGATCGGAGTTTCCAAAGAGAC 119 Desroche et al. (2005)CGTAGTAACTGCGGGAGTAATTC 102 Beltramo et al. (2006)CAGCGACTGAGCCAAATAAG 112 This workCTCTTCCGAGTCTTCAAAAGTTGAT 131 Beltramo et al. (2006)TCCATTTGCCGTTTCCTGGTTT 120 Beltramo et al. (2006)TGCCGACAACACCAACTGTTT 102 Desroche et al. (2005)

Page 3: Olguin Et Al. 2010

90 N. Olguín et al. / International Journal of Food Microbiology 144 (2010) 88–95

the analyzed time as a reference (calibrating condition). This was thestrain showing the highest ΔCt value (ΔCt=Ct studied gene — CtldhD). Therefore, the strain chosen as the reference has a RE value ofone, while the RE for the rest of the strains represents the time-fold ofover-expression for that gene, in the same conditions, compared tothe strain taken as a reference. It is worth noting that ldhD (internalcontrol) variations were less than one unit showing an average Ctclose to 21 in all sampled times and strains.

2.7. Chemical analysis

MLF was monitored by measuring L-malic acid, citric acid, aceticacid, D-glucose and D-fructose using Boehringer enzymatic kits(Mannheim). The concentration of diacetyl was detected by high-pressure liquid chromatography (HPLC Agilent 1100 Series). Themethodology used was adapted from De Revel et al. (2000) by usingthe shorter column Zorbax Eclipse XDB-C8 (150 mm×4.6 mm×5 μm) purchased from Agilent Technologies and changing theseparation and elution programmes of the mobile phase: water/acetic acetic acid 0.5%-methanol 0/20, 2/50, 6/58, 10/64, 15/75, 16/100, 18/100, 22/0, 23/20, and 28/20 (time in min/% methanol).

2.8. Enzymatic analysis

O. oeni cellular extracts were obtained by agitation on a vortex of amixture of the bacterial suspension in 10 mM of Tris–HCl buffer(pH 8.0) with glass beads. After centrifugation at 15,000 rpm for10 min, the supernatant fluid was transferred into a sterile tube andkept at −20 °C for further analysis. The malolactic (Labarre et al.,1996), citrate lyase (Cogan, 1981), acetate kinase (Cogan, 1987) andα-acetolactate decarboxylase (Ramos et al., 1995) activities wereperformed as previously described. Protein concentrations weredetermined using the Bradford (1976) method, with bovine serumalbumin as standard protein. Enzymatic activity units (U) weredefined as μmol/min.

3. Results

3.1. Strain selection

The selection of strains for this study was based on a preliminaryscreening of twelve O. oeni strains in order to choose those thatbehave differently in terms of malolactic performance and citrate use.O. oeni PSU-1was selected as a reference strain because it has been thesubject of previous studies (Olguín et al., 2009). The results of strainscreening (data not shown) were also taken into account. In terms ofmalolactic performance, PSU-1 developed a slower MLF than otherstrains. The type strain of the O. oeni species, CECT 217T, was alsoselected because it was found to have delayed MLF. Furthermore, wefound that 217T produced the highest amount of acetate among alltested strains. Strains Z42 and MO were selected because of their fastMLF performance and population viability under wine-like stressconditions. Therefore, four strains were finally selected as studymodels: strains with slow (PSU-1 and 217T) and fast (Z42 and MO)MLF, and one strain with a higher acetic acid production (217T)compared to the others.

3.2. Metabolic evolution during MLF and growth

The malolactic performance of the strains differed, with strainsZ42 and MO finishing MLF in about two days, while fermentationswith 217T and PSU-1 took five days (Fig. 1A). All strains grew duringMLF (Fig. 1B) although PSU-1 did so to a lesser extent. In all strains themaximum biomass values observed coincided with those recorded atthe end of MLF.

Citrate use was the highest in strain 217T (Fig. 1C), close to 4 mM,which represents the consumption of 87% of the initial citrate in themedium. MO also showed active citrate metabolism during MLF,consuming 58% of the citrate. However, Z42 and PSU-1 used lesscitrate, consuming only 40% and 35% of the initial citrate, respectively.The molar relation of citrate consumption and acetate production wasclose to one in MO, 217T and PSU-1, and a little bit higher (1.3) in Z42(Fig. 1C and D). The consumption of sugars was poor in general(Fig. 1E and F). Nevertheless, it is worth noting the higherconsumption of fructose by 217T between the middle and end ofMLF. Diacetyl concentrations detected in all fermentations were verylow, always showing values below the taste threshold (data notshown).

3.3. Expression of genes involved in malate and citrate metabolism

The inter-strain comparison of the relative expression (RE) ofmleA(Fig. 2) revealed a clear difference between the strains that performedfast MLF (Z42 and MO) and slow MLF (PSU-1 and 217T). Z42 and MOshowed an over-expression of 35-fold and 45-fold, respectively,compared to 217T and PSU-1, which both had an RE of close to one.Throughout MLF strains PSU-1 and 217T increased their mleAexpression compared to the inoculum (Fig. 3). Despite this increase,since transcriptional levels were initially much lower for 217T andPSU-1, these strains showed a similar mleA expression to that in Z42and MO at the end of MLF.

As shown in Fig. 2, the citrate lyase gene (citE) showed an initial 3-fold over-expression in 217T compared to the rest of the strains.However, in the middle of MLF, very low transcriptional levels of citEwere detected in 217T, while Z42 showed a considerable over-expression compared to the time of inoculation (Fig. 3). At the end ofMLF all the studied strains behaved similarly, increasing citEexpression. At this time of fermentation the maximum values of citEwere detected for all strains (lowest ΔCt values). A similar trend wasobserved in the acetate kinase gene (ackA) and acetolactatedecarboxylase (alsD). In the inocula, few differences were observedamong strains for these genes, with the exception of the over-expression of ackA in Z42 compared to the other strains (Fig. 2), andthey did not vary significantly during the course of MLF (Fig. 3).

3.4. Enzymatic activities related to MLF and citrate metabolism

According to enzymatic activity measurements (Table 2), strains217T and PSU-1 showed malolactic enzyme specific activities ofaround 30% and 20% lower than strains Z42 and MO. In the middle ofMLF the malolactic activity detected decreased in general and wascomparable for all strains.

Citrate lyase activity was similar for all strains at the time ofinoculation. However, in the middle of MLF, activity increased ingeneral and was around 30% higher in strains MO and 217T than inZ42 and PSU-1. No relevant changes were detected for acetate kinaseenzymatic activity between inoculation and mid MLF, although it isworth noting that 217T obtained lower values for this enzyme.Regarding acetolactate decarboxylase, Z42 and MO showed around30% higher enzymatic activities than PSU-1 and 217T at the time ofinoculation. This enzymatic activity did not vary in Z42 or MOthroughout MLF, whereas the activity of PSU-1 increased to similarlevels as previous strains. In the case of 217T, the activity remainedlower and even decreased a little in the middle of MLF.

3.5. Transcriptional response of stress related genes

The transcriptional behaviour of several genes related to stressresponse was compared among the four studied strains (Figs. 4 and5). These genes can be grouped according to the function that theyencode for: stress proteins (hsp18, clpP, clpX and ctsR), redox balance

Page 4: Olguin Et Al. 2010

A B

D

FE

C

0

5

10

15

20

25

30

35

40

0 1 2 3 4 5 6

days

L-m

alic

aci

d (m

M)

Citr

ic a

cid

(mM

)

Ace

tic a

cid

(mM

)

MLF(i) MLF(m) MLF(e)

Glu

cose

(m

M)

0

1

2

3

4

5

6

MLF(i) MLF(m) MLF(e)

Fru

ctos

e (m

M)

0

20

40

60

80

100

120

140

0 1 2 3 4 5 6

days

Bio

mas

s (m

g/l)

4

3

2

1

5

0

12

10

8

6

4

2

0MLF(i) MLF(m) MLF(e)

6

5

4

3

2

1

0

MLF(m)MLF(i) MLF(e)

Fig. 1. Evolution of biomass and metabolites throughout MLF in the four studied strains. Z42 (black triangles, solid line), MO (black squares, solid line), PSU-1 (empty triangles,dashed line), 217T (empty squares, dashed line). A: L-malic acid. B: biomass, dry weight. C: citric acid. D: acetic acid. E: D-glucose. F: D-fructose. Data shown are mean values withstandard deviations (n=2).

Fig. 2. Inter-strain comparison of the relative expression of the genes mleA, citE, ackA and alsD at time zero, using the strain showing the lowest ΔCT value for each gene as thereference condition.

91N. Olguín et al. / International Journal of Food Microbiology 144 (2010) 88–95

Page 5: Olguin Et Al. 2010

Fig. 3. Evolution of the relative expression ofmle, citE, ackA and alsD genes in themiddleand at the end of malolactic fermentation, using the time of inoculation as the referencecondition.

92 N. Olguín et al. / International Journal of Food Microbiology 144 (2010) 88–95

maintenance (trxA) andmembrane and cell wall biosynthesis (ddl, cfaand rmlB). They were chosen from a collection of stress related genesaccording to results obtained in a previous work (Olguín et al., 2008),in which relevant transcriptional changes of these genes wereobserved in response to ethanol in strain PSU-1.

Some common trends were observed depending on the fermen-tation stage, mainly once MLF was started. At mid MLF an increase inthe expression of some genes, such as hsp18, clpP and clpX (Fig. 4), wasobserved in strain Z42 compared to the time of inoculation. Therelevant increase in clpX resulted in similar transcriptional levels ofthis gene compared to the other strains (Fig. 5), since clpX RE in Z42was initially very low. In fact, the most relevant changes for stressrelated genes can be appreciated in the inter-strain comparison oftranscriptional levels in the middle of MLF (Fig. 5), when strains Z42and MO showed considerably higher transcriptional values than theother two strains for all the genes except for clpX and ddl, for whichsimilar values were detected for all strains. PSU-1 also showed arelative over-expression of some genes (hsp18, trxA and rmlB)compared to 217T, a strain which showed a general low transcriptionof stress responsive genes. At the end of MLF (Fig. 4), a generalinduction was observed for PSU-1 genes, compared to time zero, withthe exception of rmlB, hsp18 and trxA. On the other hand, hsp18 andtrxA showed the highest value in strain MO whereas rmlB remainedsimilar for all strains.

Table 2Mean values (n=2) of specific enzymatic activities (U/mg total protein) at time ofinoculation (inoc) and in the middle of MLF (mid).

U/mg total protein Z42 MO PSU-1 217T

Inoc Mid Inoc Mid Inoc Mid Inoc Mid

Malate decarboxylase 1.78 0.53 1.85 0.62 1.46 0.50 1.23 0.45Citrate lyase 0.11 0.32 0.09 0.48 0.04 0.29 0.04 0.42Acetate kinase 0.26 0.25 0.26 0.21 0.17 0.23 0.15 0.15Acetolactate decarboxylase 0.82 0.75 0.89 0.98 0.59 0.73 0.55 0.41

4. Discussion

This work has evaluated the different behaviour of four O. oenistrains by means of metabolic performance, the evolution ofenzymatic activities and gene expression comparison. Two strains,Z42 and MO, were chosen for their faster MLF performance, incontrast with PSU-1 and 217T, which presented a delayed MLF.Moreover, strain 217T showed a higher capacity to produce aceticacid. The fermentation assays carried out with these strains inmodified cFT80 medium reflected the different metabolic traits ofthe four strains. MLF with 217T and PSU-1 took twice as long as thatwith Z42 andMO, and the highest production of acetate was observedin 217T. Although PSU-1 and 217T showed a similar malolacticperformance, 217T grew better than PSU-1 throughout fermentation,reaching biomass values close to those observed for the faster MLFstrains. This growth might be sustained by the higher consumption ofcitrate and glucose observed during the second half of MLF for thisstrain, which would also account for the increase in volatile acidity.The co-metabolism of glucose with citrate has already been describedin O. oeni (Ramos and Santos, 1996), showing that the combination ofthe two compounds results in a higher yield of biomass than theconsumption of these compounds separately (Salou et al., 1994). Amajor consequence described of the reorientation of metabolicpathways resulting from the mixotrophic conditions is the increasein the acetate yield associated with a higher level of ATP production(Ramos et al., 1994; Salou et al., 1994). This additional ATP productionwould have provided the 217T strain with an extra source of energyfor growth, although this had no consequences on the duration ofMLF. The rest of the strains consumed less glucose than 217T.Likewise, fructose was barely consumed by any of the studied strains,including 217T.

The inter-strain comparison of the malate decarboxylase gene(mleA) at the time of inoculation yielded considerable differencesamong strains. Transcriptional levels ofmleAwere much higher in thefast MLF strains, Z42 and MO, than in the slow MLF strains, PSU-1 and217T. These relative gene expression (RE) values may correspond tohigher specific enzymatic activities, which were also higher in Z42and MO at this initial step. However, differences in enzymatic activityobserved among strains were not as acute as in gene expression.Throughout MLF transcriptional levels of mleA were equal amongstrains. This is in accordance with the similitude observed inmalolactic activities detected among strains at mid MLF, whichexperienced a general decrease. According to these results, the initialmalolactic enzyme activity in relation to its transcriptional level is thedetermining factor in MLF velocity. Coucheney et al. (2005) reportedthe malolactic enzymatic activity of freeze-dried O. oeni strains as adiscriminating parameter in evaluating their potential as malolacticstarter cultures. These authors related higher malolactic activity atlow pH (3.0) with better MLF performance in wine. The differences inspecific enzymatic activities among strains showing a distinctmalolactic performance reported by Coucheney et al. (2005) arevery close to those presented in our work. Similar results have beenobserved for O. oeni in other environments such as cider, where ahigher specific enzymatic activity of the inoculum was related tofaster MLF performance (Herrero et al., 2003).

Regarding citrate metabolism, we found an initial over-expression(3-fold) of citrate lyase gene (citE) in the inoculum of strain 217T

compared to the other strains. MO also showed slightly highertranscriptional levels of citE than Z42 and PSU-1. However, at thisinitial step, enzymatic activities did not reflect relevant differencesamong strains and were generally very low. Citrate lyase specificactivities increased during MLF and higher values were observed forMO and 217T. This increment resulted in the higher rate of citric acidconsumption by these two strains in the second half of MLF. Theactivation of citrate lyase with the addition of citrate has beendescribed in other lactic acid bacteria species (Bekal et al., 1998;

Page 6: Olguin Et Al. 2010

Fig. 4. Evolution of the relative expression of several genes related to stress response in the middle and at the end of malolactic fermentation, using the time of inoculation as thereference condition.

93N. Olguín et al. / International Journal of Food Microbiology 144 (2010) 88–95

Hugenholtz, 1993; Mellerick and Cogan, 1981). Therefore, theavailability of citric acid in the new media after inoculation mighthave activated citrate lyase activity during MLF. The evolution of citEthroughout MLF showed an early activation of the transcription of thisgene in Z42 at mid MLF, which was observed later for all strains at theend ofMLF. At that point, 217T showed the lowest RE of citE. This could

Fig. 5. Inter-strain comparison of the relative expression of genes related to stress response ireference condition. Numbers in italics on unfinished columns indicate an RE value off the

be due to the fact that citric acid had been almost depleted by thisstrain. The stronger activation of citE transcription towards the end ofMLF in the strains that barely consumed citric acid during MLF couldbe indicative of the delayed use of this substrate as a mechanism forcell survival once L-malate has been exhausted. Altogether, it seemsthat the initial over-expression of citE may be indicative of higher

n the middle of MLF, using the strain showing the lowest ΔCT value for each gene as thescale used for the graph.

Page 7: Olguin Et Al. 2010

94 N. Olguín et al. / International Journal of Food Microbiology 144 (2010) 88–95

citrate consumption during MLF. This would be in accordance withpreviously reported results (Olguín et al., 2009). These results confirmthat citric acid metabolism characteristics are strain dependent asdescribed for other LABs, such as Lactococcus (Hugenholtz, 1993).

In a previous work, acetate kinase gene (ackA) expression wasrelated to increased production of acetic acid (Olguín et al., 2009). Inthis work, the RE of ackA did not vary significantly during MLF. Theinter-strain comparison of its initial transcriptional response revealedsome differences among strains but they could not be related to thedifferences in production of acetic acid observed for the studiedstrains. This correlated with the enzymatic activities, which barelyvaried during the fermentations. Studies with the Gram-positivemodel bacterium Bacillus subtilis have shown that the expression ofackA is stimulated by the presence of glucose (Grundy et al., 1993). Inthis study, initial concentrations of sugars, including glucose, weremuch lower than in the previously reported work (Olguín et al.,2009). Observed sugar consumption and acetate production in theseassays were also much lower than that detected in the previous study.Therefore, with a low glucose concentration, similar to the sugarcontent in wine, ackA was not as active. In this case, the higherproduction of acetic acid by strain 217T is mainly associated with theuse of citric acid by citrate lyase.

The enzymatic activity of α-acetolactate decarboxylase is relatedto lower concentrations of diacetyl, since it transforms α-acetolactate(diacetyl precursor) into acetoin. No relevant differences wereobserved for the gene expression of α-acetolactate decarboxylase(alsD) either in the initial inter-strain comparison or throughout MLF.This would be in agreement with Garmyn et al. (1996), who describedthe organization of alsD and alsS genes in a single constitutivelyexpressed operon. However, previous results obtained with PSU-1(Olguín et al., 2009) revealed a transcriptional increase of alsD inresponse to lower concentrations of ethanol than those used in thiswork (10% vol/vol). This suggests that alsD transcription may bedifferently regulated depending on medium conditions, such asethanol content, pH or sugar content. Garmyn et al. (1996) detectedthe presence of two transcripts, a monocistronic unit including alsS–alsD and a single transcript of alsD. The downstream initiation at aninternal promoter may explain the presence of the distinct regulationof alsD depending on the conditions. It is worth noting that initial alsDtranscriptional levels weremuch higher than those of the other citratemetabolism studied genes in all strains (data not shown). Therefore,although the expression of this gene did not vary throughout MLF, itshigh constant transcription would account for the low diacetylconcentrations detected. This would be in accordance with themeasured α-acetolactate decarboxylase enzymatic activities, whichbarely changed and were already high at the time of inoculationcompared to the other citrate metabolism activities.

In stress related genes, a clear relationship was establishedbetween a considerably higher transcription of hsp18, clpP, ctsR,trxA, cfa and rmlB in the middle of MLF, observed in strains Z42 andMO, compared to the other strains, with their better MLF performanceand growth in wine-like conditions. The heat-shock protein Hsp18has been defined as one of the stress proteins of O. oeni, playing animportant role in the response to different environmental conditions(Beltramo et al., 2006; Guzzo et al., 1997, 1998, 2000). Furthermore,the evaluation of the abundance of Hsp18 and its gene expression hasbeen successfully used as a parameter for assessing the state ofadaptation of different strains (Coucheney et al., 2005). On the otherhand, ClpP is a stress responsive protease that can act independentlyor in association with ClpX ATPase to degrade larger specificsubstrates (Makovets et al., 1998). The different behaviour observedin the transcriptional evolution of clpP and clpX genes in this work canbe explained by the fact that clpP is regulated by ctsR, whereas clpXmay depend on other mechanisms, not yet characterized (Beltramo etal., 2004). CtsR has been reported as themaster regulator of the O. oenistress response since it is responsible for the control of the

transcription of most Clp proteins, except ClpX and Hsp18 (Grand-valet et al., 2005). At the middle of MLF, strains Z42 and MO showed aremarkably higher expression of ctsR. Also the genes under ctsRcontrol, such as hsp18 and clpP, were more induced in Z42 and MOthan in the other two strains. Although CtsR acts as a repressor of thegenes under its control, it has been suggested the existence ofadditional regulation processes (Beltramo et al. 2006). Thesealternative mechanisms could account for the simultaneous inductionof hsp18, clpP and ctsR observed. These results suggest that the inter-strain comparison of the expression of these three genes in wine-likeconditions may be indicative of a higher capacity for strain adaptationand as a result, faster MLF. Likewise, some genes related to cell walland membrane biosynthesis, mainly rmlB (dTDT-glucose-4,6-dehy-dratase), but also cfa (cyclopropane fatty acid synthase), showedrelevant over-expression in the fast MLF strains Z42 and MO. Theseresults are in agreement with those published by Da Silveira et al.(2004), who described rmlB as one of the proteins over-expressed inthe presence of ethanol in a global proteomic study. In the same way,Beltramo et al. (2006) found a relevant over-expression of cfa in wineconditions, and the thioredoxin gene (trxA) also had a higher RE in theinter-strain comparison at mid MLF in the strains Z42, MO and PSU-1.This gene is related to the maintenance of redox intracellular balanceand would be responding to the oxidative stress caused by ethanol.

The evolution of stress related genes throughout MLF revealed thelate transcriptional response of PSU-1. The over-expression of thesegenes would indicate a late recovery of this strain, which was theslowest growing of those studied. In contrast, 217T was found to havelow transcriptional levels of stress related genes throughout MLF,although it grew better than PSU-1. This strain may have sustained itssurvival and growth thanks mainly to its sugar and citrate consump-tion, with the consequent production of additional ATP and theprotection against intracellular acidification associated with citrateutilization (Martín et al., 2004).

In summary, in this study we evaluated the different behaviours offour O. oeni strains by means of measuring the evolution ofmetabolites that may have an impact on organoleptic qualities, therelated enzymatic activities and their gene expression, and thetranscriptional response of several stress responsive genes. The initialactivities of malate decarboxylase and citrate lyase would determine afaster MLF and earlier citrate consumption, respectively. The inter-strain comparison of the transcriptional levels of the correspondinggenes proved a useful tool, indicative of different metabolic traits.Similar conclusions can be drawn from the inter-strain comparison ofsome stress related genes in the middle of MLF, mainly hsp18, clpP,ctsR and rmlB. These data can be helpful for the characterisation of O.oeni strains for their use as malolactic starters. More research isneeded in order to fully understand the regulation of mechanismssuch as citrate utilization and acetate production, which candrastically change wine quality, and mechanisms leading to betterstress adaptation, which can ensure a correct MLF performance.

Acknowledgements

This work was funded by grant AGL2006-03700ALI from theSpanish Ministry of Education and Science. Nair Olguín is grateful tothe Catalan government (Generalitat de Catalunya) for a predoctoralfellowship.

References

Augagneur, Y., Ritt, J.F., Linares, D.M., Remize, F., Toudot-Maréchal, R., Garmyn, D.,Guzzo, J., 2007. Dual effect of organic acids as a function of external pH inOenococcus oeni. Archives of Microbiology 188, 147–157.

Bartowsky, E., 2005. Oenococcus oeni and malolactic fermentation — moving into themolecular arena. Australian Journal of Grape and Wine Research 11, 174–187.

Bekal, S., Diviès, C., Prévost, H., 1998. Citrate lyases of lactic acid bacteria. Le Lait 78,3–10.

Page 8: Olguin Et Al. 2010

95N. Olguín et al. / International Journal of Food Microbiology 144 (2010) 88–95

Beltramo, C., Grandvalet, C., Pierre, F., Guzzo, J., 2004. Evidence for multiple levels ofregulation of Oenococcus oeni clpP–clpL locus expression in response to stress.Journal of Bacteriology 186, 2200–2205.

Beltramo, C., Desroche, N., Tourdot-Maréchal, R., Grandvalet, C., Guzzo, J., 2006. Real-time PCR for characterizing the stress response of Oenococcus oeni in a wine-likemedium. Research in Microbiology 157, 267–274.

Bradford, M.M., 1976. A rapid and sensitive method for quantification of microgramquantities of protein utilizing the principal of protein-dye-binding. AnaliticalBiochemistry 72, 248–254.

Chomczynski, P., Sacchi, N., 1987. Single-step method of RNA isolation by acidguanidinium thiocyanate–phenol–chloroform extraction. Analytical Biochemistry162, 156–159.

Cogan, T.M., 1981. Constitutive nature of the enzymes of citrate metabolism inStreptococcus lactis subsp. diacetylactis. The Journal of Dairy Research 48, 489–495.

Cogan, T.M., 1987. Co-metabolism of citrate and glucose by Leuconostoc spp.: effects ongrowth, substrates and products. The Journal of Applied Bacteriology 63, 551–558.

Coucheney, F., Desroche, N., Bou, M., Tourdot-Maréchal, R., Dulau, L., Guzzo, J., 2005. Anew approach for selection of Oenococcus oeni strains in order to producemalolactic starters. International Journal of Food Microbiology 105, 463–470.

Cox, D.J., Henick-Kling, T., 1989. Chemiosmotic energy from malolactic fermentation.Journal of Bacteriology 171 (10), 5750–5752.

Da Silveira, M.G., Golovina, E.A., Hoekstra, F.A., Rombouts, F.M., Abee, T., 2003.Membrane fluidity adjustments in ethanol-stressed Oenococcus oeni cells. Appliedand Environmental Microbiology 69 (10), 5826–5832.

Da Silveira, M.G., Baumgärtner, M., Rombouts, F.M., Abee, T., 2004. Effect of adaptationto ethanol on cytoplasmic and membrane protein profiles of Oenococcus oeni.Applied and Environmental Microbiology 70, 2748–2755.

De Man, J.C., Rogosa, M., Sharpe, M.E., 1960. A medium for the cultivation of lactobacilli.The Journal of Applied Bacteriology 23, 130–135.

De Revel, G., Pripis-Nicolau, L., Barbe, J.C., Bertrand, A., 2000. The detection of alfa-dicarbonyl compounds in wine by formation of quinoxaline derivatives. Journal ofthe Science of Food and Agriculture 80, 102–108.

Desroche, N., Beltramo, C., Guzzo, J., 2005. Determination of an internal control to applyreverse transcription quantitative PCR to study stress response in the lactic acidbacterium Oenococcus oeni. Journal of Microbiological Methods 60, 325–333.

Garmyn, D., Monnet, C., Martineau, B., Guzzo, J., Cavin, J.F., Diviès, C., 1996. Cloning andsequencing of the gene encoding alpha-acetolactate decarboxylase from Leuco-nostoc oenos. FEMS Microbiology Letters 145, 445–450.

Grandvalet, C., Coucheney, F., Beltramo,C.,Guzzo, J., 2005.CtsR is themaster regulatorof stressresponse gene expression in Oenococcus oeni. Journal of Bacteriology 187, 5614–5623.

Grandvalet, C., Assad-García, J.S., Chu-Ky, S., Tollot, M., Guzzo, J., Gresti, J., Tourdot-Maréchal, R., 2008. Changes in membrane lipid composition in ethanol- and acid-adapted Oenococcus oeni cells: characterization of the cfa gene by heterologouscomplementation. Microbiology 154, 2611–2619.

Grundy, F.J., Waters, D.A., Allen, S.H., Henkin, T.M., 1993. Regulation of the Bacillussubtilis acetate kinase gene by CcpA. Journal of Bacteriology 175, 7348–7355.

Guzzo, J., Delmas, F., Pierre, F., Jobin, M.-P., Samyn, B., Van Beeumen, J., Cavin, J.-F.,Diviès, C., 1997. A small heat shock protein from Leuconostoc oenos induced bymultiple stresses and during stationary growth phase. Letters in AppliedMicrobiology 24, 393–396.

Guzzo, J., Jobin, M.P., Diviès, C., 1998. Increase of sulfite tolerance in Oenococcus oeni bymeans of acidic adaptation. FEMS Microbiology Letters 160, 43–47.

Guzzo, J., Jobin, M.P., Delmas, F., Fortier, L.C., Garmyn, D., Tourdot-Maréchal, R., Lee, B.,Diviès, C., 2000. Regulation of stress response in Oenococcus oeni as a function ofenvironmental changes and growth phase. International Journal of Food Microbi-ology 55, 27–31.

Herrero, M., García, L.A., Díaz, M., 2003. Malolactic bioconversion using a Oenococcusoeni strain for cider production: effect of yeast extract supplementation. Journal ofIndustrial Microbiology & Biotechnology 30, 699–704.

Hugenholtz, J., 1993. Citrate metabolism in lactic acid bacteria. FEMS MicrobiologyReviews 12, 165–178.

Jobin, M.P., Delmas, F., Garmyn, D., Diviès, C., Guzzo, J., 1997. Molecular characterizationof the gene encoding an 18-kilodalton small heat shock protein associated with themembrane of Leuconostoc oenos. Applied and Environmental Microbiology 63,609–614.

Labarre, C., Guzzo, J., Cavin, J.F., Diviès, C., 1996. Cloning and characterization of thegenes encoding the malolactic enzyme and the malate permease of Leuconostocoenos. Applied and Environmental Microbiology 62 (4), 1274–1282.

Livak, K.J., Schmittgen, T.D., 2001. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔC

T method. Methods 25, 402–408.Loubiere, P., Salou, P., Leroy, M.J., Lindley, N.D., Pareilleux, A., 1992. Electrogenic malate

uptake and improved growth energetics of the malolactic bacterium Leuconostocoenos grown on glucose–malate mixtures. Journal of Bacteriology 174, 5302–5308.

Makovets, S., Titheradge, A.J., Murria, N.E., 1998. ClpX and ClpP are essential for theefficient acquisition of genes specifying type IA and IB restriction systems.Molecular Microbiology 28, 25–35.

Martín, M.G., Sender, P.D., Peirú, S., de Mendoza, D., Magni, C., 2004. Acid-inducibletranscription of the operon encoding the citrate lyase complex of Lactococcus lactisBiovar. diacetylactis CRL264. Journal of Bacteriology 186, 5649–5660.

Mellerick, D., Cogan, T.M., 1981. Induction of some enzymes of citrate metabolism inLeuconostoc lactis and other heterofermentative lactic acid bacteria. The Journal ofDairy Research 48, 497–502.

Olguín, N., Bordons, A., Reguant, C., 2008. Expression of Oenococcus oeni genes related tocellular adaptation andmetabolic development in wine conditions. 9th Symposiumon Lactic Acid Bacteria. Egmond aan Zee, The Netherlands.

Olguín, N., Bordons, A., Reguant, C., 2009. Influence of ethanol and pH on the geneexpression of the citrate pathway in Oenococcus oeni. Food Microbiology 26,197–203.

Ramos, A., Santos, H., 1996. Citrate and sugar cofermentation in Leuconostoc oenos, a 13Cnuclear magnetic resonance study. Applied and Environmental Microbiology 62,2577–2585.

Ramos, A., Poolman, B., Santos, H., Lolkema, J.S., Konings, W.N., 1994. Uniport of anioniccitrate and proton consumption in citrate metabolism generates a proton motiveforce in Leuconostoc oenos. Journal of Bacteriology 176, 4899–4905.

Ramos, A., Lolkema, J.S., Konings, W.N., Santos, H., 1995. Enzyme basis for pH regulationof citrate and pyruvate metabolism by Leuconostoc oenos. Applied and Environ-mental Microbiology 61 (4), 1303–1310.

Salema, M., Lolkema, J.S., San Romão, M.V., Loureiro Días, M.C., 1996. The proton motiveforce generated in Leuconostoc oenos by L-malate fermentation. Journal ofBacteriology 178 (11), 3127–3132.

Salou, P., Loubiere, P., Pareilleux, A., 1994. Growth and energetics of Leuconostoc oenosduring cometabolism of glucose with citrate or fructose. Applied and Environ-mental Microbiology 60, 1459–1466.

Teixeira, H., Gonçalves, M.G., Rozès, N., Ramos, A., San Romão, M.V., 2002. Lactobacillicacid accumulation in the plasma membrane of Oenococcus oeni: a response toethanol stress? Microbial Ecology 43, 146–153.

Vaillant, H., Formisyn, P., Gerbaux, V., 1995. Malolactic fermentation of wine: study ofsome physic-chemical factors by experimental design assays. The Journal ofApplied Bacteriology 79, 640–650.

Versari, A., Parpinello, G.P., Cattaneo, M., 1999. Leuconostoc oenos and malolacticfermentation in wine: a review. Journal of Industrial Microbiology & Biotechnology23, 447–455.

Wibowo, D., Eschenbruch, R., Davis, C.R., Fleet, G.H., Lee, T.H., 1985. Occurrence andgrowth of lactic acid bacteria in wine: a review. American Journal of Enology andViticulture 36 (4), 302–313.