You are on page 1of 8

Food Hydrocolloids 17 (2003) 455462 www.elsevier.

com/locate/foodhyd

Emulsion stabilizing properties of pectin


J. Lerouxa, V. Langendorffa, G. Schickb, V. Vaishnavc, J. Mazoyera,*
a

Research Center, Degussa Texturant Systems France SAS, Baupte F50500, France b Degussa Texturant Systems GmbH, 85354, Freising, Germany c Degussa Texturant Systems Inc, Atlanta, GA 30340, USA Received 19 July 2002; revised 18 November 2002; accepted 16 December 2002

Abstract Citrus pectin and beet pectin are able to reduce the interfacial tension between an oil phase and a water phase and can be efcient for the preparation of emulsions. Investigations were made to evaluate the effect of various parameters of pectin on its emulsifying capacity. Orange and rapeseed oils emulsions were prepared with pectin as an emulsier. They were then separated by centrifugation and the pectin fraction remaining in the aqueous phase was analyzed. It was found that the molecular weight, protein and acetyl contents inuenced signicantly the emulsifying properties. It was observed that for both citrus and beet pectin, the fraction which became associated with the oil contained much more protein than the fraction in the aqueous phase. It is suggested that protein associated with the pectin played a key role in the stabilization of the emulsion. Our experiments indicated that depending on the pectin source, beet or citrus, only a limited quantity is adsorbed on the oil surface. The de-acetylated beet pectin maintained a good emulsifying ability but the chemically acetylated citrus pectin gave better results than the non-acetylated citrus pectin. It was inferred that acetyl groups could also contribute to emulsion stability. It is likely that they act by reducing the calcium bridging occulation. A model is proposed to explain the emulsifying function of pectin. q 2003 Elsevier Science Ltd. All rights reserved.
Keywords: Pectin; Sugar beet; Citrus; Emulsifying properties; Protein; Acetyl; Molecular weight

1. Introduction Pectin is a well-known food additive which is mainly used for its gelling and stabilizing abilities. It is extracted from the plant cell wall, especially citrus peels, apple pomace and sugar beet pulps. Pectin is used to make gels in aqueous media containing sugar and acid. Pectin is also able to stabilize dairy protein under acidic conditions, a role previously explained by Parker, Boulenguer and Kravtchenko (1994). The two mentioned applications account for the main worldwide consumption of pectin, but a few other functionalities have also been reported. Kertesz (1951), in an extensive review of pectin, also mentioned its emulsifying properties. As early as 1927, the use of pectin as an emulsifying agent in various applications such as avor, mineral and vegetable oils emulsions and mayonnaise, was suggested (Rooker, 1927). Pectin has a very complex structure which depends on both its source and the extraction process. Numerous studies
* Corresponding author. Tel.: 33-23-371-34-83; fax: 33-23-37134-92. E-mail address: jacques.mazoyer@degussa.com (J. Mazoyer).

contributed, and continue, to elucidate the structure of pectin. Basically, it is a polymer of a-D -galacturonic acid with 1-4 linkages (Aspinall, 1980). This ain chain is regularly interrupted by some rhamnogalacturonan segments which combine galacturonic acid residues and a-L rhamnopyranose by a 1-2 linkage (Schols & Voragen, 1996). Rhamnogalacturonan contains lateral chains which comprise of arabinan and arabinogalactan linked on O-4 or O-3 of the rhamnosyl units (Aspinall, 1980; Selvendran, 1985). The galacturonic acid of the backbone is partially methyl-esteried and O-acetylated at C-2 or C-3. In addition, lateral chains have some phenolic acids such as ferulic acid, which are linked to the arabinose and galactose via ester linkages (Fry, 1983). It is worthwhile to note that the plant primary cell wall contains proteins and particularly hydroxyproline-rich proteins (Lamport & Northcote, 1960). There is no strong evidence for any covalent linkages between pectin and glycoprotein (Ridley, ONeill, & Mohnen, 2001). However, within the analyses of various industrial pectin samples Kravtchenko, Voragen, and Pilnik (1992) have reported the presence of hydroxyproline rich protein in pectin which was

0268-005X/03/$ - see front matter q 2003 Elsevier Science Ltd. All rights reserved. doi:10.1016/S0268-005X(03)00027-4

456

J. Leroux et al. / Food Hydrocolloids 17 (2003) 455462

not completely removed by copper purication. Recently Oosterveld, Voragen, and Schols (2002) suggested that an arabinogalactan-protein was linked to the pectin extracted from hops. A comparison of the relevant chemical features of pectin from the three main sources is given in Table 1. It clearly illustrates that sugar beet is different in terms of protein and acetyl content. There is no clear explanation about the origin of the emulsifying function of pectin. Hypothesizing that a high acetyl content could enhance the hydrophobicity of pectin, Dea and Madden (1986) studied the emulsifying ability of sugar beet pectin in relation to its chemical structure. They concluded that there was no evidence for a relationship between chemical composition and emulsifying ability. Nevertheless, according to Endre and Rentschler (1999), the emulsifying ability of beet pectin can be explained by the presence of acetyl groups (4 5%). In our previous publication, Akhtar, Dickinson, Mazoyer, and Langendorff (2001), we studied the emulsifying properties of citrus pectins. This paper concluded that citrus pectin, which is low in acetyl, may have an interesting emulsifying capacity. The pectin with a low molecular weight of about 60 70 kg mol21 and a high degree of methoxylation shown the best emulsifying properties. Only a small part of the pectin which is associated with most of the protein became adsorbed onto the oil. Pectin is not the only gum to be reported with emulsifying properties. Lotskar and Maclay (1943) have found good emulsifying abilities with various gums, e.g. tragacanth, acacia, karaya and pectin. Gum arabic (Acacia senegal) is a commercially important emulsifying agent for avor oils. It is generally used at high concentrations of about 15 25% w/w in the emulsions. Its emulsifying ability is due to a small amount of protein which is covalently bound to a highly branched polysaccharide
Table 1 Galacturonic acid (GalA), rhamnose (Rha), arabinose (Ara), xylose (Xyl), galactose (Gal) and protein contents (wt%), degree of methoxylation and degree of acetylation of some acid extracted pectin Apple GalAa Rhaa Araa Xyla Gala NS (1) Protein DAc DMa
a b c d

structure (Dickinson, Elverson, & Murray, 1989; Dickinson, Galazka, & Anderson, 1991; Randall, Phillips, & Williams, 1988). In addition, other polysaccharides have been reported with emulsifying abilities, Garti and Reichman (1993) demonstrated that micro-crystallinecellulose, guar and locust bean gum were surface active, not due to protein moieties, but due to steric and mechanical stabilization mechanisms. Huang, Kakuda, and Cui (2001) reported the efciency of various hydrocolloids gums in stabilizing emulsions. The aim of this study was to compare various pectins, differing in origin and molecular weight, in terms of their emulsifying capacity and to propose a relationship between structure and emulsifying property.

2. Experimental 2.1. Materials High-molecular-weight pectins were extracted from dried citrus peels or sugar beet pulp by hydrolysis with nitric acid at pH 1.6 for 1 h at 80 8C. Depolymerized citrus pectin (DCP) samples were prepared by heating the extraction slurry at 120 8C for 10 min. A range of various molecular citrus pectin, from 13 to 186 kg mol21, were prepared according to the procedure described by Akthar, Dickinson, Mazoyer, and Langendorff (2002). After purifying the slurries by ltration, the slurry syrups were concentrated by ultraltration, and the pectin samples were recovered by precipitation in isopropyl alcohol. The products were then dried and ground. Table 2 provides characteristics of the samples. Molecular weight was measured by light scattering. The degree of methoxylation was determined by titration and the galacturonic acid content was determined by titration and colorimetry using the metahydroxydiphenyl method described by Thibault (1979). Protein content N 6:25 was determined by the Kjeldahl procedure. The acetyl content was measured according to the colorimetric dosage (McComb & McCready, 1957).
Table 2 Source, molecular weight Mw ; degree of methoxylation (DM), galacturonic acid (GalA), acetyl and protein contents (wt%) of the pectin samples. (CP: high molecular weight citrus pectin, DCP1, 2 and 3: depolymerised citrus pectin, BP1 and 2: sugar beet pectin) Samples ID Source Mw (kg/mole) 162 38 72 62 DM (%) 72.9 66.3 76.6 71.4 57.1 61.2 GalA (%) 79 81.5 83.3 80.2 79.2 81.6 Acetyl (%) 0.46 0.39 1.93 2.98 Protein (%) 0.93 21.61 1.32 0.77 1.95 2.28

Citrus 79.2 1.4 1.1 0.2 2.4 5.1 33.3b 1.41.6b 72

Beet 62.4 5.4 5.1 0.2 9.3 19.9 10.4c 16a 35d 54

73.1 2.3 4.4 1.7 4.2 12.6 1.6b 5b 74

Total neutral sugar is the sum of the mentioned sugars. Axelos and Thibault (1991). Kravtchenko, Pilnik, and Voragen (1992). Thibault (1988). Levigne, Ralet, and Thibault (2002).

CP DCP1 DCP2 DCP3 BP1 BP2

Citrus Citrus Citrus Citrus Beet Beet

J. Leroux et al. / Food Hydrocolloids 17 (2003) 455462

457

Rapeseed oil (RSO) (Bouton dor, France) was purchased in the local shops and the Bresil orange oil (OO) was provided by Degussa Food and Flavors (Grasse, France). Gum arabic was the Instant Gum AS IRX 40830 (CNI, France) and the synthetic resin was Ester Gum 8BG (Hercules BV, The Netherlands). Parafn oil was purchased from ProdHyg Laboratories, France. 2.2. Interfacial tension

The pectin was recovered from the aqueous phase by precipitation into isopropyl alcohol. The precipitate was washed in pure alcohol before drying and grinding. The proportion and composition of pectin associated with the droplets was calculated from the concentration and composition of pectin present in the aqueous phase before emulsication and that found in the serum layer after centrifugation. 2.5. Acetylation and de-acetylation procedures

The interfacial tensions were measured using the Du Nouy ring method with a tensiometer CS-Du Nouy 70535, CSC Scientic Company. The tension was measured at the interface between the parafn oil and a 2% w/w pectin solution in a pH 3.8 sodium citrate, citric acid buffer 0.02 M. 2.3. Emulsion preparation and characterization Pectin powder was added slowly to a solution containing 0.1% w/w sodium benzoate and 0.2% w/w citric acid at room temperature with gentle stirring. The pH of the resulting pectin solution was adjusted to pH 3.5 by adding 1 M NaOH. Oil-in-water emulsions (20% w/w rapeseed or orange oil) were prepared at room temperature using a laboratoryscale homogenizer ALM2 (Pierre Guerin, France) with three passes at 200 bars. The orange oil was rst mixed with Ester Gum 8BG in order to increase its density. The two phases were then mixed by a magnetic stirrer for 30 min before being homogenized. The droplet-size distributions of the emulsions were measured using a static laser lightscattering analyzer (Malvern Mastersizer 2000) equipped with liquid dispersing tank (hydro 2000S). The emulsifying ability was assessed by checking the shape of the distribution and measuring the value of the average droplet size. The average droplet size was characterized by the equivalent volume mean diameter, D4; 3; dened by: X X D4; 3 i ni di4 = i ni di3 ; where ni is the number of droplets of diameter di : This value is similar to an average volume (or weight if the density is constant) of a distribution we could have obtained by sieving. This means that only one droplet of a large size generates an increase in the mean diameter. Droplet size determination was performed after 24 h storage at room temperature and after a further 7 and 30 days in order to assess the stability of the emulsion. 2.4. Polysaccharide and protein adsorption The amount of pectin adsorbed onto the droplet surface following emulsication was inferred from measurements of the concentration of polysaccharide remaining in the serum phase after centrifugation (60,000g for 2 4 h).

The acetylation of pectin was performed according to the procedure of Carson and Maclay (1946). 5 g of commercial HM-citrus pectin (Degussa Texturant Systems, DM72.0, Acetyl cont. 0.56%) were dissolved in 150 ml of formamide. Then, 5 ml of pyridine and variable amounts of acetic anhydride (1, blank: 0 ml; 2: 1.25 ml; 3: 2.5 ml, and 4: 5.0 ml) were added and the solution was stirred for 2 h at 30 8C. The products corresponding to each amount of acetic anhydride were identied as ACP1, ACP2, ACP3 and ACP4. The acetylated pectin was precipitated, depending on the degree of substitution, with acidied methanol or acetone. Beet pectin was de-acetylated by slowly adding 3.5 ml of 50% sodium hydroxide in water to a solution of 15 g of beet pectin in 400 ml of water and stirring for 20 h at 6 8C. After careful neutralization under thorough homogenization with 7.5 ml of 25% hydrochloric acid in water, the saponied and de-acteylated beet pectin was precipitated with 1 l of isopropanol and ltered. The product was washed twice with 400 ml of isopropanol, dried and ground (Yield: 12.2 g). The resulting LM-beet pectin (DM7.4) was remethylated by dispersing the pectin powder in methanolic hydrochloric acid. Therefore, 3.6 ml of acetic chloride were added dropwise to 90 ml of methanol. After 9 g of the pectin obtained above had been added, the slurry was stirred for 24 h at 20 8C. The resulting de-acetylated HM-beet (DABP5) was recovered by ltration, washed with 70% aqueous methanol and pure methanol. Finally, the dried product was ground.

3. Results and discussion 3.1. Interfacial activity of various pectin samples First, the interfacial properties of the differing pectin samples at 2% w/w concentration were examined in comparison with gum arabic at 15% w/w concentration. Gum arabic serves as the comparison since it is the commercial emulsifying gum and is generally used at this range of concentration. The observations are shown in Table 3. The most signicant reductions of tension are observed for Depolymerised and beet pectins. There is no clear theory to explain why the low molecular weight pectins are better than those of higher molecular weights. It is likely that kinetic effects may be involved in such

458

J. Leroux et al. / Food Hydrocolloids 17 (2003) 455462

Table 3 Interfacial tensions of parafn oil/2% w/w pectin solutions at pH 3.8, at 25 8C, in (mN/m) Interfacial tension Buffer pH 3.8 Citrus pectin (CP) Citrus pectin (DCP1) Beet pulp pectin (BP1) Gum Arabica
a

36.3 31.3 20.2 19.4 19.7

Gum Arabic solution 15%.

behavior. The interfacial tensions were measured immediately after the two phases were in contact. One might suggest that high molecular products which develop more viscous solutions should move more slowly to the interface. Thus waiting for an equilibrium, might have given different results. Huang, Kakuda and Cui (2001) waited for equilibrium for a period of 30 min and observed a more signicant tension reduction for a non-depolymerised pectin. Garti and Reichman (1994) also observed this kinetic effect for more diluted guar solutions. Moreover, we should also take into account the conformational aspect of pectin which is well known to be different from gum arabic. Pectin is a semi-exible polymer whereas gum arabic adopts a random coil conformation. This may account for the surface coverage. It is interesting to note that pectin, at 2% concentration, had an effect similar to the gum arabic at 15% on the interfacial tension reduction. 3.2. Emulsifying ability of citrus and beet pectin Fig. 1, compares the particle size distributions of emulsions made with gum arabic D4; 3 0:31 mm; depolymerised citrus D4; 3 0:40 mm and citrus pectin D4; 3 0:80mm: It demonstrates that pectin is able to make emulsions in the same way as gum arabic. Nevertheless, DCP gave better results in terms of both distribution prole and mean diameter. Non-depolymerised citrus pectin showed a second peak which was attributed to the beginning of a calcium bridging occulation (Akthar, Dickinson, Mazoyer, & Langendorff, 2002). In the same study, the effect of molecular weight of citrus pectin was reported, it was established that a pectin of high DM and a molecular weight of 70 kg mol21 gave the best results in terms of particle size diameter and stability on creaming. The effect of pectin molecular weight is shown in Fig. 2 where DCPs of molecular weights between 50 and 80 kg mol21 gave the best results in terms of particle size and stability. It is noticeable that these observations are rather consistent with the reductions of the interfacial tension results. However, very low molecular weight pectin, even if it reduces the interfacial tensions, seems to lose a part of its

Fig. 1. Particle size distribution proles after 24 hours of emulsions made with orange oil 10%, 10% ester gum and 4% w/w of high molecular weight pectin ( ), of DCP ()mm and 25% of gum arabic (W).

emulsifying capacity giving coarser emulsions. It was tentatively explained in our previous paper that, since emulsions made with high molecular pectins may undergo a calcium bridging occulation, a reduction of the chain length could reduce the probability of the interactions. In comparison with the DCP, beet pectin gave better results. Fig. 3 shows a comparison between these two pectins in rapeseed oil emulsions as a function of pectin concentration. It was observed that beet pectin was very efcient for producing a ne emulsion at 2 wt%, whereas citrus pectin required higher concentrations (. 4 wt%). In terms of particle size distribution prole, beet pectin produced some Gaussian proles which were very stable on storage. Thus once again, it is shown that pectin is able to act as a food emulsier able to stabilize oil in water emulsions even

Fig. 2. Particle mean diameter D4; 3 (mm) of emulsions made with orange oil 10% and 10% ester gum in 4% w/w of various molecular weight citrus pectin solution. Measurements were made after 24 h, 7 and 20 days storage at room temperature 24 h ( ), 7 days (K) and 30 days (A).

J. Leroux et al. / Food Hydrocolloids 17 (2003) 455462

459

Fig. 3. Particle size (D4; 3 in mm) after 24 h storage at 25 8C of emulsions made with rapeseed oil at 20% in DCP (A) and beet pectin ( ) vs pectin concentration.

Fig. 5. Adsorbed pectin ( ) and adsorbed pectin fraction (A) in emulsions made with orange oil at 20% in beet pectin as a function of the pectin concentration.

those containing rather high concentrations of oil phase (20%). Pectin was good in both the avor oil and vegetable oil emulsions we studied. It is evident that beet pectin was more efcient than citrus pectin since it produced ner particle distribution prole and more stable emulsions at lower pectin concentrations. In our previous study (Akthar, Dickinson, Mazoyer, & Langendorff, 2002) on citrus pectin, we observed that the pectin fraction which became associated which the oil droplets contained almost all the protein fraction present in the hydrocolloid. Therefore, it is suggested that pectin could behave in the same way. The emulsifying properties of gum arabic are connected to a small fraction of the gum rich in protein (Randall, Phillips, & Williams, 1988).

3.3. Polysaccharide adsorption The adsorption of pectin onto the oil was studied in various media. The adsorption of the DCP onto the rapeseed oil, as a function of the pectin concentration, is shown in Fig. 4. The adsorbed pectin amount increases linearly up to about 4% pectin meaning that a constant fraction of the pectin (about 5%) is adsorbed. It also could be inferred that 4% pectin in the emulsion should correspond to the surface coverage of the droplets, this concentration will be called the adsorption threshold. However, we must be prudent since, beyond a certain concentration, the increase of viscosity makes the phase separation difcult. For high pectin concentrations some oil may remain in the serum layer. In the case of DCP with orange oil, was not possible to carry
Table 4 Weight fraction and quantity of adsorbed of pectin in various emulsions with 20% rapeseed oil (RSO) or orange oil with DCP and beet pectin. The pectin concentration corresponds to the beginning of the adsorption threshold Beet pectin orange oil Beet pectin RSO 1.5 Citrus pectin RSO 4.0

Total pectin concentration in the emulsion Weight fraction of pectin adsorbed onto the oil Quantity of pectin adsorbed onto the oil Fig. 4. Adsorbed pectin ( ) and adsorbed pectin fraction (A) in emulsions made with rapeseed oil at 20% in DCP as a function of the pectin concentration.

(w% emulsion) (w% total pectin) (mg/100 g emulsion)

2.0

9.8

14.9

4.7

196

224

188

The system of citrus pectin in orange oil is not given here because the threshold could not be observed. The higher pectin concentration required made the solution to be too much viscous to be separated correctly.

460

J. Leroux et al. / Food Hydrocolloids 17 (2003) 455462

Table 5 Weight percentage of protein and acetyl contents in initial and adsorbed pectins in various emulsions made with 1% pectin BP1/OO Whole pectin Protein Acetyl (w%) (w%) 1.95 1.93 Adsorbed fraction 21.2 3.9 BP1/RSO Whole pectin 1.95 1.93 Adsorbed fraction 7.9 2.7 DCP2/OO Whole pectin 1.32 0.46 Adsorbed fraction 13.8 2.7 DCP3/RSO Whole pectin 0.77 0.39 Adsorbed fraction 7.8 2.1

out a correct phase separation at high pectin concentrations. Therefore, for this system, no threshold could be observed. For this reason, further chemical analysis of the pectin structure will be made at lower concentrations. In the case of the beet pectin in orange oil (Fig. 5), the results are rather different. The adsorption threshold appears at a lower concentration, 2% instead of 4% for DCP. Curiously, there is an increase of adsorbed fraction as a function of the pectin concentration below the maximum, whereas it was steady with citrus pectin. The maximum of adsorption fraction is at 10%, i.e. more than with the citrus pectin. The values of the adsorption threshold for the different systems and the weight fractions of adsorbed pectin for the three systems for which we could make observations are given in Table 4. It was interestingly observed that the quantities of adsorbed pectins are rather constant whatever the system. Thus, in comparison with the citrus pectin, more beet pectin adsorbs onto the oil more and therefore less pectin is required to cover the droplet surface. These observations are consistent with the emulsion droplet size analysis. Comparisons between protein and acetyl contents in the whole initial pectin and in the adsorbed fraction are given in Table 5. This comparison is made at a pectin concentration of 1 wt% in the emulsion, it is below the maximum of adsorption. Thus, all of the active fraction of the pectin is assumed to be attached to the oil. The adsorbed pectin shows a signicant increase in both protein and acetyl content. It is observed that the composition of the adsorbed pectin, in terms of protein and acetyl contents, is more dependant on the nature of the oil than on the source of
Table 6 Particle size (D4; 3 at 24 h in mm) of emulsions made with 20% of weighted orange oil and 2 wt% pectin of chemically acetylated citrus pectins in comparison to sugar beet pectin (SBP) ACP: acetylated citrus pectin, BP sugar beet pectin Acetyl (%) ACP1 ACP2 ACP3 ACP4 BP 0.57 2.24 5.59 8.73 24 Emulsion D4; 3 (mm) 2.76 2.67 1.31 0.65 0.4 0.5

the pectin, e.g. in the rapeseed oil, even if the original pectins are different, the adsorbed fractions seems to be very similar. The pectin which reacts with the orange oil seems to have more protein, than that for the rapeseed oil. This is probably why less beet pectin than citrus pectin is required to make an emulsion with RSO. From a global point of view, as beet pectin contains more protein than citrus pectin, much less beet pectin is necessary to reach the adsorption threshold.

3.4. Acetylation and de-acetylation experiments In order to provide more information about the contribution of the acetyl groups to emulsifying functionality, two further experiments were conducted: rst, the acetylation of citrus pectin which is normally poorly acetylated and has low emulsifying properties and secondly the de-acetylation of a normally acetylated sugar beet pectin which has good emulsifying properties. The emulsifying ability of acetylated citrus pectins are presented in Table 6. This table gives the D4; 3 values after 24 h storage of the emulsions made with pectins as a function of the levels of acetylation in comparison with beet pectin. This experiment demonstrates that citrus pectin requires higher amounts of acetyl than the beet pectin to be as efcient. The de-acetylated beet pectin did not show any signicant loss in emulsifying capacity (Table 7). Therefore, although it seems the acetyl groups are more common in the pectin fraction which adheres the oil phase, their presence is not an absolute requirement with respect to the emulsifying capacity.

Table 7 Particle size after 24 h and 7 days storage D4; 324 h ; D4; 37d of emulsions made with 20% of weighted orange oil and 2 wt% pectin of chemically de-acetylated sugar beet pectins. DA-SBP 5: de-acetylated sugar beet pectin Pectin ID Acetyl (%) Protein (%) D4; 324 h (mm) D4; 37 d (mm) 1.95 2.19 0.52 0.68 0.64 0.73

SBP 1.93 DA-SBP 5 0.17

J. Leroux et al. / Food Hydrocolloids 17 (2003) 455462

461

Fig. 6. Particle size distribution proles of emulsions made with orange oil 10% and 10% ester gum in 2% w/w beet pectin (dashed line, D4; 3 mm) and de-calcied (full line, D4; 3 mm).

3.5. The effect of calcium


Fig. 7. Hypothetical model of emulsion stabilization by pectin.

The calcium content of the pectin seems to have an important effect on the emulsion stability. Akthar, Dickinson, Mazoyer, and Langendorff (2002) mentioned the likelihood that calcium would induce a bridging occulation. This calcium effect was also tested on sugar beet pectin. In this investigation, a sample of sugar beet pectin was washed in acidied isopropyl alcohol in order to lower the calcium content and the product was tested in emulsions. In this way, the calcium content was reduced from 5700 ppm down to 2060 ppm. The particle size distributions of an orange oil emulsion made with 2% pectin of both decalcied and non decalcied pectins are shown in Fig. 6. The d4.3 was reduced from 0.564 to 0.371 mm. We can observe that the distribution becomes almost perfectly Gaussian without any additional peak at about 5 mm. Thus, even with the acetylated beet pectin, the acidied alcohol washing leads to better results which are probably due to the reduction of calcium.

4. Conclusion In this study, we have shown that pectin is denitely able to produce ne and stable emulsions in the same manner as gum arabic but at much lower dosage. Among the various pectin sources, sugar beet has the best emulsifying properties. The observed emulsifying properties of pectin are most probably due to the protein residues present within the pectin. Thus the model of association to oil droplets may be similar to that of gum arabic as proposed by Randall, Phillips, and Williams (1989). However, there is

a conformational difference between pectin and gum arabic. Pectin is a semi-exible polymer whereas arbinogalactan protein complex which is the most active of part gum arabic has a coil conformation with a small radius of gyration and equivalent sphere hydrodynamic radius. Since less pectin is required to cover the oil droplet surface than gum arabic, it may be inferred that pectin takes up a greater volume around the droplets. This could be due to the more extended conformation of the pectin molecule. Pectin chains are able to strongly complex calcium and some interchain associations may arise due to calcium binding. This interaction may cause occulation. Since any acetyl groups may reduce calcium sensitivity, they also contribute to the emulsion stability avoiding the bridging occulation. Thus the combination of acetyl groups and protein is suggested to give the pectin its emulsifying properties (see suggested model in Fig. 7). The more favorable properties of sugar beet in comparison with citrus pectin may be explained by the fact that beet pectin contains more protein and more acetyl groups but it could also be due to possible conformational differences between the two pectin molecules.

Acknowledgements The authors gratefully acknowledge K. Born and the staff of the Research Center of Degussa Texturant Systems S.A.S for their help, especially, A. Bourdais and D. Callais for their kind assistance and S. Wildmoser for the preparation of

462

J. Leroux et al. / Food Hydrocolloids 17 (2003) 455462 Lamport, D. T. A., & Northcote, D. H. (1960). Hydroxyproline in primary cell walls of higher plants. Nature, 188, 665666. Levigne, S., Ralet, M.-C., & Thibault, J.-F. (2002). Characterization of pectins extracted from fresh sugar beet under different conditions using an experimental design. Carbohydrate Polymers, 49(2), 145153. Lotskar, H., & Maclay, W. D. (1943). Pectin is an emulsifying agent, comparative efciencies of pectin, tragacanth, karaya and acacia. Industrial and Engineering Chemistry, 35(12), 12941297. McComb, E. A., & McCready, R. M. (1957). Determination of acetyl in pectin and in acetylated carbohydrate polymers. Analytical Chemistry, 29(5), 819 921. Oosterveld, A., Voragen, A. G. P., & Schols, H. A. (2002). Characterization of hop pectins shows the presence of an arabinogalactan protein. Carbohydrate Polymers, 49(4), 407413. Parker, A., Boulenguer, P., & Kravtchenko, T. P. (1994). Effect of the addition of high methoxyl pectin on the rheology and colloidal stability of acid milk drinks. In K. Nishinari, & E. Doi (Eds.), Food hydrocolloids: Structure, properties and functions (pp. 307 312). New York: Plenum Press. Randall, R. C., Phillips, G. O., & Williams, P. A. (1988). The role of proteinaceous component on the emulsifying properties of gum arabic. Food Hydrocollods, 2(2), 131 140. Randall, R. C., Phillips, G. O., & Wiliams, P. A. (1989). Fractionation and characterization of gum from acacia senegal. Food Hydrocollods, 3(1), 65 75. Ridley, B. L., ONeill, M. A., & Mohnen, D. (2001). Pectins: structure, biosynthesis, and oligalacturonide-related signaling. Phytochemistry, 57, 929967. Rooker, W. A. (1927). New uses of fruit pectin. The Fruit Products Journal and American Vinegar Industry, 7(1), 11. Schols, H. A., & Voragen, A. G. J. (1996). Complex pectin: structure elucidation using enzymes. In J. Visser, & A. G. J. Voragen (Eds.), Pectins and Pectinases (Vol. 14) (pp. 319). Progress in Biotechnology, Amsterdam: Elsevier. Selvendran, R. R. (1985). Developments in the chemistry and biochemistry of pectic and hemicellulosic polymers. Journal of Cell Science, 2(Suppl.), 51 88. Thibault, J.-F. (1979). Automatisation du dosage des substances pectiques par la methode au meta-hydroxydiphenyl. Lebensmittel 2 Wissenschaft Technologie, 12, 247 251. Thibault, J.-F. (1988). Characterization and oxidative crosslinking sugarbeet pectins extracted from cossetes and pulps under different conditions. Carbohydrate Polymers, 8, 209223.

acetylated and de-acetylated pectin samples. Thanks also go to Professor E. Dickinson for helpful discussions and to Dr M Akthar.

References
Akhtar, M., Dickinson, E., Mazoyer, J., & Langendorff, V. (2002). Emulsion stabilisation of depolymerised pectin. Food Hydrocolloids, 16, 249 256. Aspinall, G. O. (1980). Chemistry of cell wall polysaccharides. The Biochemistry of Plants, New York: Academic Press. Axelos, M. A. V., & Thibault, J.-F. (1991). Inuence of the substituents of the carboxyl and of the rhamnose content on the solution properties and exibility of pectins. International Journal of Biological Macromolecules, 13, 7782. Carson, J. F., & Maclay, W. D. (1946). The acetylation of polyuronides with formamide as a dispersing agent. Journal of American Chemical Society, 68, 10151017. Dea, I. C. M., & Madden, J. K. (1986). Acetylated pectic plysaccharides of sugar beet. Food Hydrocolloids, 1(1), 7188. Dickinson, E., Elverson, D. J., & Murray, B. S. (1989). On the lm-forming and emulsion stabilising properties of gum arabic: dilution and occulation aspect. Food Hydrocolloids, 3, 101 114. Dickinson, E., Galaska, V. B., & Anderson, D. M. (1991). Emulsifying behaviour of gum arabic. Carbohydrate Polymers, 141, 373 392. Endre, H. U., & Rentschler, C. (1999). Chances and limit for the use of pectin as emulsierPart 1. The European Food and Drink Review, Summer, 4953. Fry, S. C. (1983). Feruloylated pectins from the primary cell wall: their structure and possible functions. Planta, 157, 11 123. Garti, N., & Reichman, D. (1993). Hydrocolloids as food emulsiers and stabilizers. Food Structure, 12, 411 426. Garti, N., & Reichman, D. (1994). Surface properties and emulsication activity of galactomannans. Food Hydrocolloids, 8(2), 155173. Huang, X., Kakuda, Y., & Cui, W. (2001). Hydrocolloids in emulsions: particle size distribution and interfacial activity. Food Hydrocolloids, 15, 533 542. Kertesz, Z. I. (1951). The pectic substances. New York: Interscience Publishers, Inc. Kravtchenko, T. P., Voragen, A. G. J., & Pilnik, W. (1992). Analytical comparison of three industrial pectin preparations. Carbohydrate Polymers, 18(1), 17 25.

You might also like