Homopolysaccharides from lactic acid bacteria

11
International Dairy Journal 11 (2001) 675–685 Homopolysaccharides from lactic acid bacteria Pierre Monsan*, Sophie Bozonnet, C! ecile Albenne, Gilles Joucla, Ren ! e-Marc Willemot, Magali Remaud-Sim! eon Centre de Bioing ! enierie Gilbert Durand, DGBA-INSA, UMR CNRS 5504, UMR INRA 792, INSA, De ´pt de Ge ´nie Biochimique and Alimentaire, 135 avenue de Rangueil, 310077 Toulouse Cedex 4, France Abstract In addition to heteropolysaccharides of complex structure, lactic bacteria produce a variety of homopolysaccharides containing only either d-fructose or d-glucose. These fructans and glucans have a common feature in being synthesized by extracellular transglycosylases (glycansucrases) using sucrose as glycosyl donor. The energy of the osidic bond of sucrose enables the efficient transfer of a d-fructosyl or d-glucosyl residue via the formation of a covalent glycosyl-enzyme intermediate. In addition to the synthesis of high molecular weight homopolysaccharides, glycansucrases generally catalyse the synthesis of low molecular weight oligosaccharides or glycoconjugates when efficient acceptors, like maltose, are added to the reaction medium. While the enzymatic synthesis of fructans (levan and inulin) is poorly documented at the molecular level, the field of Streptococcus and Leuconostoc glucansucrases (glucosyltransferases and dextransucrases) has been well studied, both at the mechanistic and gene structure levels. The nutritional applications of the corresponding polysaccharides and oligosaccharides account for this increasing interest. r 2001 Elsevier Science Ltd. All rights reserved. Keywords: Fructan; Levan; Inulin; Glucan; Dextran; Mutan; Alternan; Fructansucrase; Glucansucrase; Glucosyltransferase 1. Introduction Lactic acid bacteria produce a wide variety of exopolysaccharides (van Geel-Schutten, 2000), which are mainly involved in cell adhesion and protection. Until recently, industrial interest has resulted from their physical–chemical properties, but these polysaccharides have now raised new interest due to their potential for nutritional and health applications. In addition to heteropolysaccharides composed of glucose, galactose, fructose and rhamnose (De Vuyst & Degeest, 1999), lactic acid bacteria produce homopoly- saccharides which contain only one type of monosac- charide, fructose or glucose, respectively, the fructans and the glucans. In fact, most of these homopolysac- charides share the feature of being synthesized by extracellular glycansucrases using sucrose as the glycosyl (fructose or glucose) donor. Unlike the majority of polysaccharides, these poly- mers are not generally produced by glycosyltransferases which use nucleotide-sugar precursors, but by transgly- cosylases (glycansucrases) which are able to use the energy of the osidic bond of sucrose to catalyse the transfer of a corresponding glycosyl moiety: In addition to the synthesis of high-molecular-mass polymers, glycansucrases generally catalyse the produc- tion of low-molecular-mass oligosaccharides when efficient acceptor molecules, such as maltose, are present in the reaction mixture in addition to sucrose (Koepsell et al., 1952): *Corresponding author. Tel.: +33-561-55-94-15; fax: +33-561-55- 94-00. E-mail address: [email protected] (P. Monsan) 0958-6946/01/$ - see front matter r 2001 Elsevier Science Ltd. All rights reserved. PII:S0958-6946(01)00113-3

Transcript of Homopolysaccharides from lactic acid bacteria

International Dairy Journal 11 (2001) 675–685

Homopolysaccharides from lactic acid bacteria

Pierre Monsan*, Sophie Bozonnet, C!ecile Albenne, Gilles Joucla,Ren!e-Marc Willemot, Magali Remaud-Sim!eon

Centre de Bioing !enierie Gilbert Durand, DGBA-INSA, UMR CNRS 5504, UMR INRA 792, INSA, Dept de Genie Biochimique and Alimentaire,

135 avenue de Rangueil, 310077 Toulouse Cedex 4, France

Abstract

In addition to heteropolysaccharides of complex structure, lactic bacteria produce a variety of homopolysaccharides containing

only either d-fructose or d-glucose. These fructans and glucans have a common feature in being synthesized by extracellulartransglycosylases (glycansucrases) using sucrose as glycosyl donor. The energy of the osidic bond of sucrose enables the efficienttransfer of a d-fructosyl or d-glucosyl residue via the formation of a covalent glycosyl-enzyme intermediate. In addition to thesynthesis of high molecular weight homopolysaccharides, glycansucrases generally catalyse the synthesis of low molecular weight

oligosaccharides or glycoconjugates when efficient acceptors, like maltose, are added to the reaction medium. While the enzymaticsynthesis of fructans (levan and inulin) is poorly documented at the molecular level, the field of Streptococcus and Leuconostocglucansucrases (glucosyltransferases and dextransucrases) has been well studied, both at the mechanistic and gene structure levels.

The nutritional applications of the corresponding polysaccharides and oligosaccharides account for this increasing interest. r 2001Elsevier Science Ltd. All rights reserved.

Keywords: Fructan; Levan; Inulin; Glucan; Dextran; Mutan; Alternan; Fructansucrase; Glucansucrase; Glucosyltransferase

1. Introduction

Lactic acid bacteria produce a wide variety ofexopolysaccharides (van Geel-Schutten, 2000), whichare mainly involved in cell adhesion and protection.Until recently, industrial interest has resulted from theirphysical–chemical properties, but these polysaccharideshave now raised new interest due to their potential fornutritional and health applications.In addition to heteropolysaccharides composed of

glucose, galactose, fructose and rhamnose (De Vuyst &Degeest, 1999), lactic acid bacteria produce homopoly-saccharides which contain only one type of monosac-charide, fructose or glucose, respectively, the fructansand the glucans. In fact, most of these homopolysac-charides share the feature of being synthesized byextracellular glycansucrases using sucrose as the glycosyl(fructose or glucose) donor.

Unlike the majority of polysaccharides, these poly-mers are not generally produced by glycosyltransferaseswhich use nucleotide-sugar precursors, but by transgly-cosylases (glycansucrases) which are able to use theenergy of the osidic bond of sucrose to catalyse thetransfer of a corresponding glycosyl moiety:

In addition to the synthesis of high-molecular-masspolymers, glycansucrases generally catalyse the produc-tion of low-molecular-mass oligosaccharides whenefficient acceptor molecules, such as maltose, are presentin the reaction mixture in addition to sucrose (Koepsellet al., 1952):

*Corresponding author. Tel.: +33-561-55-94-15; fax: +33-561-55-

94-00.

E-mail address: [email protected] (P. Monsan)

0958-6946/01/$ - see front matter r 2001 Elsevier Science Ltd. All rights reserved.

PII: S 0 9 5 8 - 6 9 4 6 ( 0 1 ) 0 0 1 1 3 - 3

2. Fructans

Two types of fructose homopolysaccharides areproduced by fructosyltransferases from sucrose: levanand inulin, which contain b-2,6 and b-2,1 osidic bonds,respectively (Fig. 1).Little is known about fructan-synthesizing sucrases.

The only reported 3D structure is that of the levansu-crase from Bacillus subtilis, but at a poor level ofresolution, 0.38 nm (Lebrun & van Rapenbusch, 1980).The proposed mechanism of catalysis for fructosyl-transferases is a two-step mechanism involving bifunc-tional catalysis in which an acidic group and anucleophilic group of the enzymes are involved in thetransfructosylation reaction (Sinnot, 1990).

2.1. Levan

Levansucrase (E.C. 2.4.1.10) catalyses the transfer ofd-fructosyl residues from fructose to yield the b-2,6osidic bonds which characterize levan (Fig. 1). Thesynthesis of this fructan has been studied in most detailin B. subtilis and Zymomonas mobilis, but no significantapplications have as yet been developed involving thispolysaccharide.Among lactic acid bacteria, levansucrase is produced

by strains from the oral flora, such as Streptococcussalivarius and Streptococcus mutans (Giffard, Allen,Milward, Simpson, & Jacques, 1993; Shiroza & Kur-amitsu, 1988). The enzyme from S. salivarius has amolecular mass of 140 kDa. It is bound to the cell wall,

but is partly released into the culture medium in thepresence of sucrose (Milward & Jacques, 1990).Levansucrase is produced in cultures of Leuconostoc

mesenteroides NRRL B-512F in addition to dextran-sucrase. Levansucrase is responsible for at least 25%of the reducing sugars produced when grown in thepresence of sucrose (Robyt & Walseth, 1979). This is asignificant level of glucose and explains why thereis fructose repression of dextransucrase production inLeuc. mesenteroides NRRL B-512F and demonstrationof levansucrase activity (Dols, Remaud-Sim!eon, &Monsan, 1998).The presence of a levansucrase activity in Lactoba-

cillus reuteri LB 121 was also recently reported (VanGeel-Schutten et al., 1999; van Geel-Schutten, 2000). Itwas also associated with a glucansucrase activity, butchemostat cultures of this strain resulted in rapidaccumulation of spontaneous exopolysaccharide-nega-tive mutants without any sucrase activity. Mutants withno levansucrase activity emerged following a pH shift-down (van Geel-Schutten, 2000).

2.2. Inulin-type

Fructooligosaccharides containing b-2,1 osidic bonds(Fig. 1) are of nutritional interest, as they are non-digestible and potentially present very interesting pre-biotic properties for both humans and animals (Toku-naga, Nakada, Tashiro, Hirayama, & Hidaka, 1993;Bouhnik et al., 1999; Diplock et al., 1999). They areobtained either by enzymatic synthesis from sucrose,using a fungal fructosyltransferase, or by controlledhydrolysis of inulin polymers. But, as yet, suchfructooligosaccharides have not been synthesized usingfructosyltransferases from lactic acid bacteria, eventhough this group of organisms has this type ofenzymatic activity.

S. mutans Ingbritt A strain produces a fructan whichonly contains b-2,1 linked fructosyl units (Baird, Long-year, & Ellwood, 1973), while S. mutans JC-2 producesan inulin-type fructan consisting mainly of b-2,1linked fructosyl units with 5% b-2,6 branches(Rosell & Birkhed, 1974; Ebisu, Kato, Kotani, &Misaki, 1975).Very surprisingly, L. reuteri LB 121 which is known to

produce a linear levan (see above), also contains a ftfAgene (2400 bp) which was isolated with PCR techniques(van Geel-Schutten, 2000). In the presence of sucrose,the corresponding fructosyltransferase, FTFA, producesfructooligosaccharides with a degree of polymerisationof 3–4, and a high molecular mass polysaccharide withb-2,1 linked fructosyl units only. FTFA contains aputative N-terminal secretion signal peptide. It shows90% similarity with SACB of S. mutans, which alsosynthesizes a fructan containing only b-2,1 osidic bonds.Fig. 1. Structures of fructans: (A) levan, (B) inulin-type fructan.

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685676

3. Glucans

Glucansucrases catalyse the synthesis of a variety ofglucans containing mostly a-1,6, a-1,3, a-1,4 and a-1,2linked d-glucosyl units (Sidebotham, 1974; Monchois,Willemot, & Monsan, 1999). The structure (Fig. 2) ofthese glucans has been elucidated using a wide range ofmethods: periodic oxidation, polymer methylation,acetolysis, Smith degradation, enzymatic hydrolysisand 13C NMR (Jeanes et al., 1954; Seymour, Knapp,& Bishop, 1976; Seymour, Slodki, Plattner, & Jeanes,

1977; Seymour & Knapp, 1980; Misaki, Torii, Savai, &Goldstein, 1980; C #ot!e & Robyt, 1982a).Extracellular glucansucrases are mostly produced by

lactic acid bacteria belonging to the genera: Leucono-stoc, Streptococcus and Lactobacillus (Sidebotham,1974; Mooser, 1992). The only exception is the produc-tion of amylosucrase (E.C. 2.4.1.4), which catalyses thesynthesis of a-1,4-d-glucosyl linkages from sucrose, byNeisseria sp. (MacKenzie, McDonald, & Johnson, 1978;Potocki de Montalk, Remaud-Sim!eon, Willemot, Plan-chot, & Monsan, 1999). Amylosucrase is the only

Fig. 2. Structures of glucans: (A) dextran, (B) mutan and, (C) alternan.

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685 677

glucansucrase belonging to family 13 of glycosidehydrolases, while all the other glucansucrases belongto family 70 (Henrissat & Davies, 1997).Various strains produce more than one glucansucrase:

three distinct enzymes are produced by S. mutans 6715(serotype g) (Shimamura, Tsumori, & Mukasa, 1983),four are produced by S. sobrinus (Walker, Cheetham,Taylor, Pearce, & Slodki, 1990), three by L. mesente-roides NRRL B-1355 (Smith, Zahnley, Wong, Lundin,& Ahlgren, 1998).The glucansucrases from Streptococcus sp. are gen-

erally produced constitutively, whereas those fromLeuconostoc sp. are specifically induced by sucrose,except for some constitutive mutants (Mizutani, Yama-da, Takayama, & Shoda, 1994; Kim & Robyt, 1995).Streptococcal glucansucrases (generally named gluco-

syltransferases), particularly from S. mutans, are in-volved in cariogenesis phenomena. In fact, the glucanpolymers synthesized by these enzymes play a key role inthe adhesion of bacteria to the tooth surface to formdental plaque (Muzaka & Slade, 1973; Hamada & Slade,1979; Hamada & Slade, 1980).Despite a large amount of scientific work, the

catalytic mechanism of glucansucrases has still not beentotally elucidated. The key central step of the transfer ofthe d-glucosyl unit is the formation of a covalentglucosyl-enzyme intermediate. This step involves, as inthe case of the enzymes of family 13 of glucosidehydrolases (Henrissat & Davies, 1997), in particulara-amylases, a catalytic triad consisting of two asparticacids and one glutamic acid residue (MacGregor,Jesperen, & Svensson, 1996; Devulapalle, Goodman,Gao, Hemsley, & Mooser, 1997). Such a catalytic sitewas clearly identified by Mooser and Iwakoa (1989) whoisolated a covalent glucosyl-enzyme complex from aquenched reaction of the glucosyltransferase from S.sobrinus with radiolabelled sucrose. An aspartic acidresidue was identified at the catalytic site of two S.sobrinus glucosyltransferases (Mooser, Hefta, Paxton, &Lee, 1991).From this covalent glucosyl-enzyme intermediate, the

glucosyl residue can be transferred into a series ofacceptors (Koepsell et al., 1952; Ebert & Schenk, 1968)as follows:

* to the growing dextran chain to extend it by onecarbohydrate unit;

* to a carbohydrate or a non-carbohydrate acceptor toyield an oligosaccharide or a gluco-conjugate;

* to a water molecule: this corresponds to the hydro-lysis of the sucrose molecule (Robyt & Corrrigan,1977; Robyt & Walseth, 1979; Luzio, Parnaik, &Mayer, 1983; Yokoyama, Kobayashi, & Matsuda,1985);

* to a fructose molecule to give either neo-synthesis of asucrose molecule, i.e. ‘‘isotopic exchange’’ (Mayer,

Matthews, Futerman, Parnaik, & Jung, 1981; Jung &Mayer, 1981) or the production of leucrose: a-d-glucopyranosyl-1,5-d-fructopyranose (Stodola,Koepsell, & Sharpe, 1956; Schwengers, 1991).

In addition, in the absence of residual sucrose in thereaction medium, glucansucrases can catalyse dispro-portionation reactions involving oligosaccharides assubstrates (Binder, C #ot!e, & Robyt, 1983; Lopez-Munguia et al., 1993).Glucan polymer synthesis follows a processive me-

chanism. This is deduced from the observations thatintermediate oligosaccharides cannot be detected in thereaction medium during the synthesis, and high mole-cular weight polysaccharides are obtained at earlyreaction times ( Tsuchiya, Hellman, & Koepsell, 1953;Bovey, 1959; Ebert et al., 1968).Two alternative mechanisms have been proposed for

the glucan chain growth (Monchois et al., 1999):

* Non-reducing end elongation: Only one covalentglucosyl-enzyme intermediate is involved (Mooser,1992). In the case of dextransucrase, the initialacceptor could be the dextran when associated tothe enzyme (Hehre, 1941), but polymer synthesisoccurs even in the absence of dextran (Ebert &Schenk, 1968; Robyt & Corrrigan, 1977; Miller,Ecklund, & Robyt, 1986). Sucrose has been suggestedas the initial acceptor by Neely (1960), in disagree-ment with the results of Parnaik and Mayer (1982)and Su and Robyt (1994). But Cheethman, Slodki,and Walker (1991) have shown that a sucrose residueoccurs at the end of the dextran chain produced bythe glucosyltransferase GTF-S3 from S. sobrinus.

* Reducing end elongation: This mechanism was sug-gested by Ebert and Schenk (1968) and demonstratedby Robyt, Kimble, and Walseth (1974) in the case ofimmobilized dextransucrase from Leuc. mesenteroidesNRRL B-512F using a pulse and chase reaction with14C-labelled sucrose. Similar results have beenobtained with the glycosyltransferases from S. mutans6715 (Robyt & Martin, 1983) and S. sanguis ATCC10558 (Ditson & Mayer, 1984). This mechanisminvolves two identical nucleophilic sites able to yieldtwo covalent glucosyl-enzyme intermediates fromtwo sucrose molecules (Robyt et al., 1974). The C-6hydroxyl from one of these two glucosyl residuesmakes a nucleophilic attack on the C-1 of the secondglucosyl intermediate to give an a-1,6 glucosidiclinkage between the two glucosyl moieties. Thereleased nucleophilic site then attacks another glucosemolecule to give a new glucosyl-enzyme intermediate.This symmetrical and alternative role of the twonucleophilic sites results in the growth of the glucanchain by its reducing end, without the need for aprimer and without the release of the glucan chainfrom the enzyme before the addition of the next

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685678

glucosyl residue. Branching of the glucan chain doesnot necessarily involve any additional enzyme, butcan occur when such a chain acts as an acceptor ofglucosyl units.

It is difficult to make a clear choice between these twomechanisms, but until now sequence analysis andstructure prediction of known glucansucrases has onlyresulted in the identification of a single active site(MacGregor et al., 1996; Devulapalle et al., 1997).In the case of the acceptor reaction, glucosyl units are

clearly added at the non-reducing end of the acceptoroligosaccharides (Koepsell et al., 1952; Robyt &Walseth, 1978). Robyt and Martin (1983) have sug-gested an oligosaccharide synthesis mechanism invol-ving independently the two nucleophilic sitesparticipating in the polysaccharide synthesis. But thesuppression of one of the potential nucleophilic sites bysite-directed mutagenesis, i.e. the replacement of aspar-tic acid residue 551 of Leuc. mesenteroides NRRL B-512F dextransucrase by an asparagine residue, totallysuppresses the ability of the enzyme to catalyse thesynthesis of oligosaccharides in the presence of maltoseas acceptor (Monchois, Remaud-Sim!eon, Russell, Mon-san, & Willemot, 1997). This result is not compatiblewith an identical role for the two nucleophilic sites.More than thirty genes encoding glucansucrases have

been isolated and sequenced. The enzymes are closelyrelated and share a common structure (Fig. 3). They arecomposed of four distinct structural domains (Monchoiset al., 1999):

* The N-terminal domain begins with a well conservedsignal peptide (Monchois, Remaud-Sim!eon, Monsan,& Willemot, 1998a), as these enzymes are extracel-lular. The only exception is the dextransucrase geneDSR-A from Leuc. mesenteroides NRRL B-1299(Monchois, Willemot, Remaud-Sim!eon, Croux, &Monsan, 1996).

* A highly variable stretch of 123–129 amino acids,which does not seem to play an important role inenzyme activity (Abo et al., 1991), and is not presentin DSR-A from Leuc. mesenteroides NRRL B-1299(Monchois et al., 1996).

* A highly conserved core region of about 1000 aminoacid containing the active site, and more particularlythe catalytic triad, as demonstrated for the glucosyl-transferases GTF-I and GTF-S from S. sobrinus

(Mooser & Iwakoa, 1989; Mooser, Hefta, Paxton, &Lee, 1991). This domain presents a permutated (b=a)8barrel structure, when compared with family 13 of theglycoside hydrolases (Monchois et al., 1999).

* A C-terminal domain composed of a series ofrepeating units (Ferretti, Gilpin, & Russell, 1987;Gilmore, Russell, & Ferretti, 1990; Abo et al., 1991),which is responsible for the binding of the synthesizedglucan (Monchois et al., 1999). The presence of theC-terminal domain is generally necessary to keep anenzyme active. In the case of Leuc. mesenteroidesNRRL B-512F dextransucrase (DSR-S), C-terminaldomain truncation results in a strong decrease ofactivity, both for polysaccharide and oligosaccharidesynthesis (Monchois, Reverte, Remaud-Sim!eon,Monsan, & Willemot, 1998b).

Even if the structural characterization of an increas-ing number of glucansucrase encoding genes, coupledwith a site-directed mutagenesis approach, allows abetter insight into the role of the various parts of theglucansucrases, additional research is still necessary toreach a full understanding of the catalytic mechanismsof both polysaccharide and oligosaccharide synthesis.Obviously, the elucidation of the 3D structure of thistype of enzyme will be a key step in such an under-standing, as well as in the possibility to controlglucansucrase selectivity. In fact, it is already possibleto modify the structure of the products by site-directedmutagenesis of the glucansucrases: the presence of acarboxylic amino acid instead of a threonine at position589 of the amino acid sequence of a glucansucrase fromS. mutans GS-5 synthesising an a-1,6 linked glucan(GTF-S), results in a 30% increase in the synthesis of a-1,3 glucosidic bonds (Shimamura, Nakano, Musaka, &Kuramitsu, 1994). Similar results have been obtainedwith the dextransucrase from Leuc. mesenteroidesNRRL B-512F, DSR-S: when the threonine residue atposition 667 is replaced by an arginine, the dextranobtained contains 13% of a-1,3 glucosidic bonds insteadof 5% with the native enzyme (Remaud-Sim!eon, Will-emot, Sar-cabal, Potocki de Montalk, & Monsan, 2000).Several types of glucans are obtained from the action

of glucansucrase:

3.1. Dextran

Pasteur (1861) discovered the microbial origin of thegelification of cane sugar syrups. In 1874, the corre-sponding product was named ‘‘dextran’’, due to itspositive rotatory power. The microorganism responsibleof the gelification was isolated by Van Tieghem (1878)and named it Leuc. mesenteroides. In 1941, Hehredemonstrated that dextran could be synthesized fromsucrose by a cell-free filtrate. The corresponding

Fig. 3. Schematic general structure of glucansucrases, showing: (A) N-

terminal signal-sequence, (B) variable region, (C) catalytic domain

and, (D) glucan binding domain (GBD).

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685 679

extracellular enzyme was named ‘‘dextransucrase’’ byHestrin, Averini-Shapiro, and Aschner (1943).Dextransucrase (E.C. 2.4.1.5) produces a glucan

(Fig. 2) which contains at least 50% of a-1,6 osidicbonds within the main chain (Buchholz & Monsan,2001). The degree of branching, involving a-1,2, a-1,3and a-1,4 linkages in dextrans A, B and C, respectively,(Seymour & Knapp, 1980) varies according to the originof the dextransucrase (Table 1).The most widely used dextran is produced by the

dextransucrase of the strain Leuc. mesenteroides NRRLB-512F, which synthesizes a very linear polysaccharidecontaining 95% a-1,6 linkages. The controlled chemicalhydrolysis of this high-molecular-mass dextran allowsthe production of fractions with an average molecularmass of 70 kDa. These products are used for theproduction of chromatography supports for gel permea-tion separation (Sephadexs), in the production of bloodplasma substitutes (Groenwall & Ingelman, 1948), andto prepare dextran sulphate for blood coagulationprevention and iron transport (Soetaert, Schwengers,Buchholz, & Vandamme, 1995).The molecular mass of the corresponding dextransu-

crase has been the subject of discussion. Values rangingfrom 65 kDa (Kobayashi & Matsuda, 1980) to 190 kDa(Willemot, Monsan, & Durand, 1988) have beenreported. Finally, the isolation of the gene encodingthis enzyme (Wilke-Douglas, Perchorowicz, Houck, &Thomas, 1989) led to a protein with a total of 1527amino acids (MM: 170 kDa). Lower values are due toproteolytic degradation, while higher values can beattributed to dextran contamination of the proteinpreparation.

Lb. reuteri LB 121 cells growing on sucrose synthe-size, besides a fructan polymer, large amounts of aglucan polymer with a molecular mass of 3500 kDa.This glucan has a unique structure consisting of

terminal, 4-substituted, 6-substituted, and 4,6-disubsti-tuted a-glucose in a molar ratio of 1.1 : 2.7 : 1.5 : 1.0 (VanGeel-Schutten et al., 1999). This homopolysaccharide issynthesized from sucrose by a glucansucrase (MM:146 kDa). The corresponding gene (gtfA) has beenisolated and sequenced. It shows a high homology withthe known glucansucrase genes from lactic bacteria,particularly with the gene encoding alternansucrase(ASR) from Leuc. mesenteroides NRRL B-1355 (VanGeel-Schutten, 2000). The glucansucrase ORF has a sizeof approximately 4100 bp (Van Geel-Schutten, 2000).Mutant strains lacking levansucrase activity, but keep-ing dextransucrase activity, were obtained following apH shiftdown (Van Geel-Schutten et al., 1999).The maltose acceptor reaction is used to produce non-

digestible gluco-oligosaccharides (Fig. 4) with the dex-transucrase from Leuc. mesenteroides NRRL B-1299,which is known to catalyse the synthesis of dextranpolymers containing a-1,2 linked branched chains(Table 1). Dextransucrase keeps its selectivity in theacceptor reaction and produces a series of gluco-oligosaccharides containing such a-1,2 glucosidic lin-kages (Paul, Lopez-Munguia, Remaud, Pelenc, &Monsan, 1992; Remaud-Sim!eon, Lopez-Munguia, Pe-lenc, Paul, & Monsan, 1994; Dols, Remaud-Sim!eon,Willemot, & Vignon, 1998), which resist the attack ofdigestive enzymes (Valette et al., 1993). Such gluco-oligosaccharides present interesting prebiotic properties(Monsan & Paul, 1995; Djouzy et al., 1995). Thesegluco-oligosaccharides are presently marketed for hu-man nutritional and dermocosmetic applications (BioE-colias: BioEurope/Solabia).

3.2. Mutan

Mutansucrase (E.C. 2.4.1.5) produces a water-in-soluble glucan containing more than 50% of a-1,3

Table 1

Structure of the different glucans produced by Leuconostoc mesenteroides and Streptococcus sp. glucansucrases

Osidic linkages

a-1,6 (%) a-1,3 (%) a-1,4 (%) a-1,2 (%)

Leuconostoc mesenteroides B-512F 95 5

B-742 87 13

50 50

B-1355 95 5

54 46

B-1299 66 7 27

65 35

Streptococcus downei Mfe28 12 88

90 10

Streptococcus mutans GS5 13 87

15 85

70 30

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685680

glucosidic linkages (Fig. 2), mainly associated with a-1,6linkages. This enzyme is produced by Leuc. mesente-roides NRRL B-523, B-1149 and several Streptococcusstrains (Sidebotham, 1974; Mooser, 1992).Mutan polysaccharides are involved in the adhesion

of oral flora microorganisms on the tooth surface toform the dental plaque (Hamada & Slade, 1980). Nospecific applications of mutan polymers have beendeveloped up till now.

3.3. Alternan

Alternansucrase (E.C. 2.4.1.140) synthesizes the glu-can, alternan, which contains alternating a-1,6 and a-1,3

glucosidic linkages, with some degree of a-1,3 branch-ings. Three Leuc. mesenteroides strains are known toproduce alternansucrase: Leuc. mesenteroides NRRL B-1355, NRRL B-1501 and NRRL B-1498 (Jeanes et al.,1954; Seymour & Knapp, 1980; C #ot!e & Robyt, 1982a).Jeanes et al. (1954) were the first to report the

production of two different glucansucrases by Leuc.mesenteroides NRRL B-1355. Ethanol precipitation hasresulted in the isolation of two polysaccharide fractions:(i) a poorly water-soluble polysaccharide (Fraction L)and (ii) a water-soluble polysaccharide (Fraction S).Fraction L corresponds to a polysaccha-ride containing95% of a-1,6 glucosidic linkages and 5% a-1,3 branch-ing, similar to the dextran produced by the dextransu-

Fig. 4. Maltose acceptor reactions catalysed by Leuc. mesenteroides B-1299 dextransucrase with examples of the different families of gluco-

oligosaccharide products: OD4, containing only a-1,6 osidic bonds (in addition to the a-1,4 bond of the maltose moiety at the reducing end); R5 with

a terminal a-1,2 glucosidic linkage; and R06 with a-1,2 glucosidic branching on the penultimate glucosyl residue of the non-reducing end.

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685 681

crase from Leuc. mesenteroidesNRRL B-512F. FractionS contains alternan and is produced by alternansucrase.This latter enzyme activity is mainly (85%) bound to thebacterial cells and is more thermostable than thedextransucrase activity. The association of enzyme tocells allows efficient removal of the contaminatingdextransucrase activity to give a purified alternansucrasepreparation (Lopez-Munguia et al., 1993).In the presence of sucrose and maltose acting as an

acceptor, alternansucrase retains its selectivity, andcatalyses the synthesis of oligosaccharides containingalternating a-1,6 and a-1,3 linkages (Koepsell et al.,1952; C #ot!e & Robyt, 1982b; Pelenc et al., 1991).Mutant strains of Leuc. mesenteroides NRRL B-

1355C producing alternansucrase on a glucose medium,have been obtained by Kim and Robyt (1994), Smith,Zahnley, and Goodman (1994) and by Smith andZahnley (1997). This has resulted in the identificationof a third glucansucrase in this strain, which synthesizesa water-insoluble dextran-type polymer (Zahnley &Smith, 1995).The gene encoding the alternansucrase (ASR) from

Leuc. mesenteroides NRRL B-1355 has been cloned,sequenced and expressed in E. coli (Arg .uello-Moraleset al., 2000). It contains 6171 bp and encodes a proteinof 2057 amino acids (MM: 229 kDa), which correspondsto the largest glucansucrase described to date. This geneproduces an enzyme with a structural organizationsimilar to that of known glucansucrases from lactic acidbacteria. Alternansucrase thus, belongs to family 70 ofglycoside hydrolases (Henrissat & Davies, 1997) as dothe other glucansucrases from Leuc. mesenteroides. Thevariable region and the C-terminal glucan-bindingdomain are longer than in other glucansucrases, 100and 200 amino acids, respectively. The catalytic domainpresents 49% identity with that of the other glucansu-crases from Leuc. mesenteroides. The purified enzymehas been characterized for its ability to synthesizeoligosaccharides from acceptor carbohydrates: its capa-city to use cellobiose as an acceptor is significantlyhigher than that of the dextransucrase from Leuc.mesenteroides NRRL B-512F (Arg .uello-Morales, Re-maud-Sim!eon, Willemot, Vignon, & Monsan, 2001).Such gluco-oligosaccharides are of potential interest asprebiotics.

3.4. b-1,3 glucan

Lactobacillus subsp. G-77 has been reported toproduce two glucose homopolysaccharides when grownon a glucose medium (Duenas-Chasco et al., 1998). Oneof the exopolysaccharides was shown to be a 2-substituted-(1–3)-b-d-glucan identical to that describedfor the exopolysaccharide from Pediococcus damnosus2.6 (Duenas-Chasco et al., 1997). This is the first reportof the production of a b-glucan by lactic bacteria. The

mechanism of synthesis has not as yet been describedbut it does not involve any glucansucrase, becausesucrose, which is the enzyme inducer and the glucosyldonor, was not present in the culture medium. A similarpolysaccharide is produced by P. damnosus strainsisolated from spoiled wine (Walling, Gindreau, &Lonvaud-Funel, 2001). The second homopolysaccharideis a dextran-type polysaccharide (a-1,6 glucosyl lin-kages) with a-1,2 branching of a single d-glucose unitbut its mechanism of synthesis has not been reported.These polysaccharides have undesirable properties for

some industries as they cause thickening and are at theorigin of a cider (Duenas-Chasco et al., 1997) or a wine(Gindreau, Walling, & Lonvaud-Funel, 2001) spoilageproblem known as ‘‘oiliness’’ or ‘‘ropiness’’.

References

Abo, H., Matsumura, T., Kodama, T., Ohta, H., Fukui, K., Kato, K.,

& Kagawa, H. (1991). Peptide sequences for sucrose splitting and

glucan binding within Streptococcus sobrinus glucosyltransferase

(Water-insoluble glucan synthetase). Journal of Bacteriology, 173,

989–996.

Arg .uello-Morales, M., Remaud-Sim!eon, M., Willemot, R.-M., Vig-

non, M. R., & Monsan, P. (2001). Novel oligosaccharides

synthesised from cellobiose by alternansucrase acceptor reaction.

Carbohydrate Research, 331, 403–411.

Arg .uello-Morales, M. A., Remaud-Sim!eon, M., Pizzut, S., Sar-cabal,

P., Willemot, R. M., & Monsan, P. (2000). Sequence analysis of the

gene encoding alternansucrase, a sucrose glucosyltransferase from

Leuconostoc mesenteroides NRRL B-1355. FEMS Microbiology

Letters, 182, 81–85.

Baird, J. K., Longyear, V. M. C., & Ellwood, D. C. (1973). Water

insoluble and soluble glucans produced by extracellular glycosyl-

transferases from Streptococcus mutans. Microbios, 8, 143–150.

Binder, T., C #ot!e, G., & Robyt, J. F. (1983). Disproportionation

reactions catalysed by Leuconostoc and Streptococcus glucansu-

crases. Carbohydrate Research, 124, 275–286.

Bouhnik, Y., Vahedi, K., Achour, L., Salfati, J., Pochart, P., Marteau,

P., Flouri!e, B., Bornet, F., & Rambaud, J. C. (1999). Dose-related

effect of short-chain fructo-oligosaccharides administration on the

colonic Bifidobacteria in healthy humans. Journal of Nutrition, 129,

113–116.

Bovey, F. (1959). Enzymatic polymerisation I. Molecular weight and

branching during the formation of dextran. Journal of Polymer

Science, 35, 167–182.

Buchholz, K., & Monsan, P. (2001). Dextransucrases. In: J. R.

Whitaker (Ed.), Handbook of food enzymology. New York:

Marcel Dekker, in press.

Cheethman, N., Slodki, M., & Walker, G. (1991). Structure of the

linear, low molecular weight dextran synthesized by a D-

glucosyltransferase (GTF-S3) of Streptococcus sobrinus. Carbohy-

drate Polymers, 16, 341–353.

C #ot!e, G., & Robyt, J. F. (1982a). Isolation and partial characterization

of an extracellular glucansucrase from Leuconostoc mesenteroides

NRRL B-1355 that synthesizes an alterning (1–6), (1–3)-a-D-Glucan. Carbohydrate Research, 101, 57–74.

C #ot!e, G., & Robyt, J. F. (1982b). Acceptor reactions of alternansu-

crase from Leuconostoc mesenteroides NRRL B-1355. Carbohy-

drate Research, 101, 127–142.

De Vuyst, L., & Degeest, B. (1999). Heteropolysaccharides from lactic

acid bacteria. FEMS Microbiological Reviews, 23, 153–177.

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685682

Devulapalle, K., Goodman, S., Gao, Q., Hemsley, A., & Mooser, G.

(1997). Knowledge-based model of a glucosyltransferase from oral

bacterial group of mutant streptococci. Protein Science, 6,

2489–2493.

Diplock, A. T., Aggett, P. J., Ashwell, M., Bornet, F., Fern, E. B., &

Roberfroid, M. B. (1999). Scientific concepts of functional foods in

Europe: Consensus document. British Journal of Nutrition,

81(Suppl. 1), S1–S27.

Ditson, S., & Mayer, R. (1984). Dextransucrase. The direction of chain

growth during autopolymerisation. Carbohydrate Research, 126,

170–175.

Djouzy, Z., Andrieux, C., Pelenc, V., Somarriba, S., Popot, F., Paul,

F., Monsan, P., & Szylit, O. (1995). Degradation and fermentation

of a-gluco-oligosaccharides by bacterial strains from human colon:

In vitro and in vivo studies in gnotobiotic rats. Journal of Applied

Bacteriology, 79, 117–127.

Dols, M., Remaud-Sim!eon, M., & Monsan P. (1998). Optimization of

the production of dextransucrase from Leuconostoc mesenteroides

NRRL B-1299 and its application to the synthesis of non digestible

glucooligosaccharides. In: S. F. de Azevedo, E. C. Ferreira, K. Ch.

A. M. Luyben, & P. Osseweijer (Eds.), Proceedings second

European symposium on biochemical science (pp. 86–92), Porto:

European Federation of Biotechnology, Section on Biochemical

Engineering Science.

Dols, M., Remaud-Sim!eon, M., Willemot, R. M., Vignon, M., &

Monsan, P. (1998). Structural characterization of the maltose

acceptor products synthesized by Leuconostoc mesenteroides

NRRL B-1299 dextransucrase. Carbohydrate Research, 305,

549–559.

Duenas-Chasco, M. T., Rodriguez-Carvajal, M. A., Tejero-Mateo, P.,

Espartero, J. L., Irastorza-Iribas, A., & Gil-Serrano, A. M. (1998).

Structural analysis of the exopolysaccharides produced by Lacto-

bacillus spp. G-77. Carbohydrate Research, 307, 125–133.

Duenas-Chasco, M. T., Rodriguez-Carvajal, M. A., Tejero-Mateo, P.,

Franco-Rodriguez, G., Espartero, J. L., Irastorza-Iribas, A., & Gil-

Serrano, A. M. (1997). Structural analysis of the exopolysaccharide

produced by Pediococcus damnosus 2.6. Carbohydrate Research,

303, 453–458.

Ebert, K., & Schenk, G. (1968). Mechanism of biopolymer growth:

The formation of dextran and levan. Advances in Enzymology, 30,

179–210.

Ebisu, S., Kato, K., Kotani, S., & Misaki, A. (1975). Structural

differences in fructans elaborated by Streptococcus mutans and

Strep. salivarius. Journal of Biochemistry, 78, 879–887.

Ferretti, J. J., Gilpin, M. L., & Russell, R. R. B. (1987). Nucleotide

sequence of a glucosyltransferase gene from Streptococcus sobrinus

Mfe28. Journal of Bacteriology, 169, 4271–4278.

Giffard, P. M., Allen, D. M., Milward, C. P., Simpson, C. L., &

Jacques, N. A. (1993). Sequence of the gtfK gene of Streptococcus

salivarius ATCC 25975 and evolution of the gtf genes in

streptococci. Journal of General Microbiology, 139, 1511–1522.

Gilmore, K. S., Russell, R. R. B., & Ferretti, J. J. (1990). Analysis of

the Streptococcus downei gtfS gene, which specifies a glucosyl-

transferase that synthesizes soluble glucans. Infection and Immu-

nity, 58, 2452–2458.

Gindreau, E., Walling, E., & Lonvaud-Funel, A. (2001). Direct

polymerase chain reaction detection of ropy Pediococcus

damnosus strains in wine. Journal of Applied Microbiology, 90,

535–542.

Groenwall, A. J., & Ingelman, B. G. A. (1948).Manufacture of infusion

and injection fluids. US Patent, 2,437,518.

Hamada, S., & Slade, H. D. (1979). Synthesis and binding of

glucosyltransferase and in vitro adherence of Streptococcus mutans

grown in a synthetic medium. Archives of Oral Biology, 24,

399–402.

Hamada, S., & Slade, H. D. (1980). Biology, immunology, and

cariogenicity of Streptococcus mutans. Microbiological Reviews, 44,

331–384.

Hehre, E. J. (1941). Production from sucrose of a serologically

reactive polysaccharide by a sterile bacterial extract. Science, 93,

237–238.

Henrissat, B., & Davies, G. (1997). Structural and sequence based

classification of glycoside hydrolases. Current Opinions in Structu-

ral Biology, 7, 637–644.

Hestrin, S., Averini-Shapiro, S., & Aschner, M. (1943). The enzymic

production of levan. Biochemistry Journal, 37, 450–456.

Jeanes, A., Haynes, W. C., Williams, C. A., Rankin, J. C., Melvin,

E. H., Austin, M. J., Cluskey, J. E., Fisher, B. E., Tsuchiya, H. M.,

& Rist, C. E. (1954). Characterization and classification of dextrans

from ninety six strains of bacteria. Journal of the American

Chemical Society, 76, 5041–5052.

Jung, M., & Mayer, R. (1981). Dextransucrase: Donor substrate

reactions. Archives in Biochemistry and Biophysics, 208, 288–295.

Kim, D., & Robyt, J. (1995). Production, selection and characteris-

tics of mutants of Leuconostoc mesenteroides B-742 constitutive

for dextransucrase. Enzyme and Microbial Technolology, 17,

689–695.

Kim, D., & Robyt, J. F. (1994). Production and selection of mutants of

Leuconostoc mesenteroides constitutive for glucansucrases. Enzyme

and Microbial Technology, 16, 659–664.

Kobayashi, M., & Matsuda, K. (1980). Characterization of the

multiple forms and main components of dextransucrase from

Leuconostoc mesenteroides NRRL B-512F. Biochimica Biophysica

Acta, 614, 46–62.

Koepsell, H., Tsuchiya, H., Hellman, N., Kasenko, A., Hoffman, C.,

Sharpe, E., & Jackson, R. (1952). Enzymatic synthesis of dextran.

Acceptor specificity and chain initiation. Journal of Biological

Chemistry, 200, 793–801.

Lebrun, E., & Van Rapenbusch, R. (1980). The structure of Bacillus

subtilis levansucrase at 3.8 (A resolution. Journal of Biological

Chemistry, 255, 12034–12036.

Lopez-Munguia, A., Pelenc, V., Remaud, M., Biton, J., Michel, J.,

Lang, C., & Monsan, P. (1993). Production and purification of

alternansucrase, glucosyltransferase from Leuconostoc mesente-

roides NRRL B-1355 for the synthesis of oligoalternans. Enzyme

and Microbial Technology, 15, 77–85.

Luzio, G., Parnaik, V., & Mayer, R. (1983). D-glucosylated form of

dextransucrase: Demonstration of partial reactions. Carbohydrate

Research, 121, 269–278.

MacGregor, A., Jesperen, H., & Svensson, B. (1996). Circularly

permutated a-amylase-type a/b-barrel structure in glucan-synthe-

sizing glucosyltransferases. FEBS Letters, 398, 263–266.

MacKenzie, C. R., McDonald, I. J., & Johnson, K. G. (1978).

Glycogen metabolism in the genus Neisseria: Synthesis from

sucrose by amylosucrase. Canadian Journal of Microbiology, 24,

357–362.

Mayer, R., Matthews, M., Futerman, C., Parnaik, V., & Jung, S.

(1981). Dextransucrase: Acceptor substrate reactions. Archives in

Biochemistry and Biophysics, 208, 278–287.

Miller, A. W., Ecklund, S. H., & Robyt, J. (1986). Milligram to gram

scale purification and characterization of dextransucrase from

Leuconostoc mesenteroides NRRL B-512F. Carbohydrate Research,

147, 119–133.

Milward, C. P., & Jacques, N. A. (1990). Secretion of fructosyl-

transferase by Streptococcus salivarius involves the sucrose-

dependent release of the cell-bound form. Journal of General

Microbiology, 136, 165–169.

Misaki, A., Torii, M., Savai, T., & Goldstein, I. (1980). Structure of

the dextran of Leuconostoc mesenteroides NRRL B-1355. Carbo-

hydrate Research, 84, 273–285.

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685 683

Mizutani, N., Yamada, M., Takayama, K., & Shoda, M. (1994).

Constitutive mutants for dextransucrase from Leuconostoc mesen-

teroides NRRL B-512F. Journal of Fermentation Bioengineering,

77, 248–251.

Monchois, V., Remaud-Sim!eon, M., Monsan, P., & Willemot, R. M.

(1998a). Cloning and sequencing of an extracellular dextransucrase

(DSR-B) from Leuconostoc mesenteroides NRRL B1299-synthesiz-

ing only a(1–6) glucan. FEMS Microbiology Letters, 159, 307–315.

Monchois, V., Remaud-Sim!eon, M., Russell, R. R. B., Monsan, P., &

Willemot, R. M. (1997). Characterization of Leuconostoc mesente-

roidesNRRL B-512F dextransucrase (DSR-S) and identification of

amino acid residues playing a key role in enzyme activity. Applied

Microbiology and Biotechnology, 48, 465–472.

Monchois, V., Reverte, A., Remaud-Sim!eon, M., Monsan, P., &

Willemot, R. M. (1998b). Effect of Leuconostoc mesenteroides

NRRL B-512F dextransucrase carboxy-terminal deletions on

dextran and oligosaccaride synthesis. Applied and Environmental

Microbiology, 64, 1644–1649.

Monchois, V., Willemot, R. M., & Monsan, P. (1999). Glucansucrases:

Mechanism of action and structure–function relationships. FEMS

Microbiology Reviews, 23, 131–151.

Monchois, V., Willemot, R. M., Remaud-Sim!eon, M., Croux, C., &

Monsan, P. (1996). Cloning and sequencing of a gene coding for a

novel dextransucrase from Leuconostoc mesenteroides NRRL

B-1299 synthesizing only a(1–6) and a(1–3) linkages. Gene, 182,23–32.

Monsan, P. F., & Paul, F. (1995). Oligosaccharide feed additives. In

R. J. Wallace, & A. Chesson (Eds.), Biotechnology in animal feeds

and feeding (pp. 233–245). Weinheim: VCH.

Mooser, G. (1992). Glycosidases and glycosyltransferases. The

Enzymes, 20, 187–221.

Mooser, G., Hefta, S. A., Paxton, R. J., Shively, J. E., & Lee, T. D.

(1991). Isolation and sequence of an active-site peptide containing a

catalytic aspartic acid from two Streptococcus sobrinus–glucosul-

transferases. Journal of Biological Chemistry, 266, 8916–8922.

Mooser, G., & Iwakoa, K. R. (1989). Sucrose 6-a-D-glucosyltran-sferase from Streptococcus sobrinus: Characterisation of a glucosyl-

enzyme complex. Biochemistry, 28, 443–449.

Muzaka, H., & Slade, H. D. (1973). Mechanism of adherence of

Streptococcus mutans to smooth surfaces I. Roles of insoluble

dextran-levan synthetase enzymes and cell wall polysaccharide

antigen in plaque formation. Infection and Immunity, 8, 555–562.

Neely, W. (1960). Dextran: Structure and synthesis. Advances in

Carbohydrate Chemistry, 15, 341–369.

Parnaik, V., & Mayer, R. (1982). Mechanism of chain initiation by

dextransucrase. Absence of terminal sucrose linkage in newly

synthesized dextran chains. FEBS Letters, 150, 482–484.

Pasteur, L. (1861). Sur la fermentation visqueuse. Bulletin de la Soci !et !e

de Chimie, 30–31.

Paul, F., Lopez-Munguia, A., Remaud, M., Pelenc, V., & Monsan, P.

(1992). Process for the enzymatic preparation of oligodextrans useful

in the production of sugar substitutes, and these oligodextrans. US

Patent 5,141,858.

Pelenc, V., Lopez-Munguia, A., Remaud, M., Biton, J., Michel, J.,

Paul, F., & Monsan, P. (1991). Enzymatic synthesis of oligoalter-

nans. Sciences des Aliments, 11, 465–476.

Potocki de Montalk, G., Remaud-Sim!eon, M., Willemot, R. M.,

Planchot, V., & Monsan, P. (1999). Sequence analysis of the gene

encoding amylosucrase from Neisseria polysaccharea and cha-

racterization of the recombinant enzyme. Journal of Bacteriology,

181, 375–381.

Remaud-Sim!eon, M., Lopez-Munguia, A., Pelenc, V., Paul, F., &

Monsan, P. (1994). Production and use of glucosyltransferases

from Leuconostoc mesenteroides NRRL B-1299 for the synthesis of

oligosaccharides containing a(1–2) linkages. Applied Biochemistry

and Biotechnology, 44, 101–117.

Remaud-Sim!eon, M., Willemot, R. M., Sar-cabal, P., Potocki de

Montalk, G., & Monsan, P. (2000). Glucansucrases: Molecular

engineering and oligosaccharide synthesis. Journal of Molecular

Catalysis B. Enzymatic, 10, 117–128.

Robyt, J. F., & Corrrigan, A. (1977). The mechanism of dextransu-

crase action. Archives in Biochemistry and Biophysics, 183, 726–731.

Robyt, J. F., Kimble, B., & Walseth, T. (1974). The mechanism of

dextransucrase action: Direction of chain growth during dextran

biosynthesis. Archives in Biochemistry and Biophysics, 165,

634–640.

Robyt, J. F., & Martin, P. (1983). Mechanism of synthesis of D-

glucans by D-glucosyltransferases from Streptococcus mutans 6715.

Carbohydrate Research, 113, 301–315.

Robyt, J. F., & Walseth, T. (1978). The mechanism of acceptor

reactions of Leuconostoc mesenteroides NRRL B-512F dextransu-

crase. Carbohydrate Research, 61, 433–435.

Robyt, J. F., & Walseth, T. (1979). Production, purification and

properties of dextransucrases from Leuconostoc mesenteroides

NRRL B-512F. Carbohydrate Research, 68, 95–111.

Rosell, K. G., & Birkhed, D. (1974). An inulin-like fructan produced

by Streptococcus mutans, strain JC2. Acta Chimica Scandinavica,

28, 589.

Schwengers, D. (1991). Leucrose a ketodisaccharide of industrial

design. In F. W. Lichtenthaler (Ed.), Carbohydrates as Organic

Raw Materials (pp. 183–195). Weinheim: VCH.

Seymour, F. R., & Knapp, R. D. (1980). Structural analysis of

dextrans from strains of Leuconostoc and related genera, that

contain 3-O-a glucosylated-D-glucopyranosyl residues at the

branched points or in consecutive linear position. Carbohydrate

Research, 81, 105–129.

Seymour, F. R., Knapp, R. D., & Bishop, S. H. (1976). Determination

of the structure of dextran by C-nuclear resonance spectrometry.

Carbohydrate Research, 51, 179–194.

Seymour, F. R., Slodki, M., Plattner, R., & Jeanes, A. (1977). Six

unusual dextran methylation structural analysis by combined

GLC-MS of per-O-acetyl-aldonitriles. Carbohydrate Research, 53,

153–166.

Shimamura, A., Nakano, Y. J., Musaka, H., & Kuramitsu, H. K.

(1994). Identification of amino acid residues in Streptococcus

mutans glucosyltransferases influencing the structure of the glucan

product. Journal of Bacteriology, 176, 4845–4850.

Shimamura, A., Tsumori, H., & Mukasa, H. (1983). Three kinds of

extracellular glucosyltransferases from Streptococcus mutans 6715

(serotype g). FEBS Letters, 157, 79–84.

Shiroza, T., & Kuramitsu, H. K. (1988). Sequence analysis of the

Streptococcus mutans fructosyltransferase gene and flanking

regions. Journal of Bacteriology, 175, 810–816.

Sidebotham, R. (1974). Dextrans. Advances in Carbohydrate Chemistry

and Biochemistry, 30, 371–444.

Sinnot, M. L. (1990). Catalytic mechanisms of enzymatic glycosyl

transfer. Chemical Reviews, 90, 1171–1202.

Smith, M. R., & Zahnley, J. (1997). Leuconostoc mesenteroides B-1355

mutants producing alternansucrase exhibiting decreases in appa-

rent molecular mass. Applied and Environmental Microbiology, 63,

581–586.

Smith, M. R., Zahnley, J., & Goodman, N. (1994). Glucosyltrans-

ferases mutants of Leuconostoc mesenteroides NRRL B-1355.

Applied and Environmental Microbiology, 60, 2723–2731.

Smith, M. R., Zahnley, J. C., Wong, R. Y., Lundin, R. E., & Ahlgren,

J. A. (1998). A mutant strain of Leuconostoc mesenteroides B-1355

producing a glucosyltransferase synthesizing a(1,2) glucosidic

linkages. Journal of Industrial Microbiology and Biotechnology,

21, 37–45.

Soetaert, W., Schwengers, D., Buchholz, K., & Vandamme, E. J.

(1995). A wide range of carbohydrate modifications by a single

microorganism: Leuconostoc mesenteroides. In S. B. Petersen, B.

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685684

Svensson, & S. Pedersen, Carbohydrate Bioengineering. Progress in

Biotechnology, Vol. 10 (pp. 351–358). Amsterdam: Elsevier.

Stodola, F. H., Koepsell, H. J., & Sharpe, E. S. (1956). The

preparation, properties and structure of the disaccharide leucrose.

Journal of the American Chemical Society, 78, 2514–2518.

Su, D., & Robyt, J. F. (1994). Determination of the number of sucrose

and acceptor binding sites for Leuconostoc mesenteroides B-512m

dextransucrase and the confirmation of the two site mechanism for

dextran synthesis. Archives in Biochemistry and Biophysics, 308,

471–476.

Tokunaga, T., Nakada, Y., Tashiro, Y., Hirayama, M., & Hidaka,

H. (1993). Effects of fructo-oligosaccharides on the intestinal

microflora and defecation in healthy volunteers. Bifidus, 6,

143–150.

Tsuchiya, H., Hellman, N., & Koepsell, H. (1953). Factors affecting

molecular weight of enzymatically synthesized dextran. Journal of

the American Chemical Society, 75, 757–758.

Valette, P., Pelenc, V., Djouzi, Z., Andrieux, C., Paul, F., Monsan, P.,

& Szylit, O. (1993). Bioavailability of new synthesized glucooligo-

saccharides in the intestinal tract of gnotobiotic rats. Journal of

Sciences of Food and Agriculture, 62, 121–127.

Van Geel-Schutten, G. H. (2000). Exopolysaccharide synthesis by

Lactobacillus reuteri. Ph.D. Dissertation, University of Groningen,

The Netherlands.

Van Geel-Schutten, G. H., Faber, E. J., Smit, E., Bonting, K., Smith,

M. R., Ten Brink, B., Kamerling, J. P., Vliegenthart, J. F. G., &

Dijkhuizen, L. (1999). Biochemical and structural characterization

of the glucan and fructan exopolysaccharides synthesized by the

Lactobacillus reuteri wild-type strain and by mutant strains.

Applied and Environmental Microbiology, 65, 3008–3014.

Van Tieghem, P. (1878). On sugar-mill gum. Annales des Sciences

Naturelles Botaniques. Biologiques et V !eg !etales, 7, 180–203.

Walker, G. J., Cheetham, N. W. H., Taylor, C., Pearce, B. J., & Slodki,

M. E. (1990). Productivity of four a-D-glucosyltransferasesreleased by Streptococcus sobrinus under defined conditions in

continuous culture. Carbohydrate Polymers, 13, 399–421.

Walling, E., Gindreau, E., & Lonvaud-Funel, A. (2001). La

biosynth"ese d’exopolysaccharides par des souches de Pediococcus

damnosus isol!ees du vin: Mise au point d’outils mol!eculaires de

d!etection. Lait, 81, 289–300.

Wilke-Douglas, M., Perchorowicz, J. T., Houck, C. M., & Thomas, B.

R. (1989). Methods and compositions for altering physical char-

acteristics of fruit and fruit products. PCT Patent WO 89/12386.

Willemot, R. M., Monsan, P., & Durand, G. (1988). Effects of dextran

on the activity and stability of dextransucrase from Leuconostoc

mesenteroides. Annals of the New York Academy of Sciences, 542,

169–172.

Yokoyama, I., Kobayashi, M., & Matsuda, K. (1985). Comparison of

the multiplicity of dextransucrases from six strains of Leuconostoc

mesenteroides. Agricultural and Biological Chemistry, 61,

1120–1123.

Zahnley, J., & Smith, M. R. (1995). Insoluble glucan formation by

Leuconostoc mesenteroides NRRL B-1355. Applied and Environ-

mental Microbiology, 61, 1120–1123.

P. Monsan et al. / International Dairy Journal 11 (2001) 675–685 685