You are on page 1of 11

Article

pubs.acs.org/JAFC

Interaction of a Dietary Fiber (Pectin) with Gastrointestinal


Components (Bile Salts, Calcium, and Lipase): A Calorimetry,
Electrophoresis, and Turbidity Study
Mauricio Espinal-Ruiz,†,‡ Fabián Parada-Alfonso,† Luz-Patricia Restrepo-Sánchez,†
Carlos-Eduardo Narváez-Cuenca,† and David Julian McClements*,‡,§

Departamento de Quı ́mica, Facultad de Ciencias, Universidad Nacional de Colombia, AA 14490 Bogotá, Colombia

Food Biopolymers and Colloids Research Laboratory, Department of Food Science, University of Massachusetts, 102 Holdsworth
Way, Amherst, Massachusetts 01003, United States
§
Department of Biochemistry, Faculty of Science, King Abdulaziz University, P.O. Box 80203, Jeddah 21589, Saudi Arabia

ABSTRACT: An in vitro gastrointestinal model consisting of oral, gastric, and intestinal phases was used to elucidate the impact
of pectin on the digestion of emulsified lipids. Pectin reduced the extent of lipid digestion, which was attributed to its binding
interactions with specific gastrointestinal components. The interaction of pectin with bile salts, lipase, CaCl2, and NaCl was
therefore investigated by turbidity, microstructure, electrophoresis, and isothermal titration calorimetry (ITC) at pH 7.0 and 37
°C. ITC showed that the interaction of pectin was endothermic with bile salts, but exothermic with CaCl2, NaCl, and lipase.
Electrophoresis, microstructure, and turbidity measurements showed that anionic pectin formed electrostatic complexes with
calcium ions, which may have decreased lipid digestion due to increased lipid flocculation or microgel formation because this
would reduce the surface area of lipid exposed to the lipase. This research provides valuable insights into the physicochemical and
molecular mechanisms of the interaction of pectin with gastrointestinal components that may affect the rate and extent of lipid
digestion.
KEYWORDS: pectin, gastrointestinal tract, lipid digestion, isothermal titration calorimetry, turbidity, flocculation

■ INTRODUCTION
There has been an appreciable increase in the total amount of
arabinose), linked together. The GalA units have carboxyl
groups, which may be present as free carboxyl groups or
calories consumed by humans during the past few decades, methyl-esterified groups depending on the orign, isolation, and
which is believed to be an important contributing factor to processing of pectin.13−17 Pectin has been successfully used for
increases in obesity, diabetes, and cardiovascular diseases.1 Fat many years in the food and beverage industries as a thickening
has the highest calorie density of the major food components and gelling agent, as well as a colloidal stabilizer.13,18 The
(fat, protein, and carbohydrates) and so there has been a major gelling characteristics of pectin have been utilized to form
focus on the identification of effective strategies to reduce the hydrogel delivery systems for a range of pharmaceutical and
fat content of foods while maintaining their desirable quality food bioactive compounds.18,19
attributes. Several studies have suggested that certain types of Previous studies have shown that pectin reduces the rate and
dietary fibers can inhibit lipid digestion and absorption in the extent of lipid digestion by interacting with various food and
small intestine.2−9 Increased consumption of dietary fiber may intestinal components.3,6,8 A number of different physicochem-
therefore be one approach for reducing some of the adverse ical and physiological mechanisms have been proposed to
affects associated with eating high-fat food products. account for this effect. Pectin can interact with bile salts and
Dietary fibers can be classified as either water-soluble or phospholipids in the small intestine, which may alter lipid
water-insoluble.10 Water-insoluble fibers, such as lignins, digestion by reducing the amount of surface-active components
celluloses, and some hemicelluloses, play an important role in available to stabilize the lipid droplets (triacylglycerols) or to
regulating intestinal peristalsis.11 Water-soluble fibers, such as solubilize and transport lipid digestion products (free fatty acids
pectin, carageenan, xanthan gum, and alginate, influence the and monoacylglycerols) from the droplet surfaces to the
gastrointestinal fate of ingested foods due to their ability to epithelium cells.7,9 The binding of bile salts to pectin in the
bind water, thicken or gel intestinal fluids, and interact with small intestine has also been proposed as one of the major
specific food and gastrointestinal components.12 Pectin is a mechanisms responsible for the ability of pectin to reduce
water-soluble dietary fiber that is used widely in the cholesterol levels.20 Pectin molecules may interfere with the
pharmaceutical, biotechnology, and food industries as a reabsorption of bile salts in the small intestine, thereby reducing
functional ingredient.13 Pectin is a polysaccharide that has
linear anionic regions formed by D-galacturonic acid (GalA) Received: October 6, 2014
monomers, linked by α-(1,4) glycosidic bonds, and branched Revised: November 17, 2014
regions primarily formed by various types of neutral Accepted: December 5, 2014
monosaccharides (mainly rhamnose, xylose, mannose, and Published: December 5, 2014

© 2014 American Chemical Society 12620 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630
Journal of Agricultural and Food Chemistry Article

the amount of cholesterol absorbed and transported to the triacetin) for 30 min of incubation (1 unit of lipase activity was defined
blood.20 Furthermore, the conformation and aggregation state as the amount of enzyme required for the release of 1 μequiv of fatty
of pectin in aqueous solutions may be altered due to its acid from either triacetin (pH 7.4) or olive oil (pH 7.7) in 1 h at 37
interactions with certain gastrointestinal components, which °C). The composition of the bile extract was reported as 49% (w/w)
total bile salt (BS), containing 10−15% glycodeoxycholic acid, 3−9%
leads to changes in solution rheology that might affect lipid taurodeoxycholic acid, 0.5−7% deoxycholic acid, 1−5% hydro-
digestion, for example, by altering gastric emptying times or the deoxycholic acid, and 0.5−2% cholic acid; 5% (w/w) phosphatidylcho-
mass transport of digestive enzymes.3 Pectin may also directly line (PC); Ca2+ ≤ 0.06% (w/w); critical micelle concentration of bile
interact with other components in the gastrointestinal tract extract 0.07 ± 0.04 mM; and mole ratio of BS to PC being around
such as CaCl2, NaCl, and digestive enzymes.9,21−23 Finally, 15:1. All other chemicals were purchased from Sigma-Aldrich
pectin is known to promote depletion flocculation of lipid Chemical Co. Double-distilled water was used to prepare all solutions.
droplets, which may reduce lipid digestion by protecting them Simulated Gastrointestinal Studies. Solution and Emulsion
from attack by lipolytic enzymes.24,25 Preparation. Pectin stock solution (4% w/w) was prepared by
dispersing 4 g of powdered pectin into 96 g of 5 mM phosphate buffer
Pectin ingredients obtained from different natural sources solution (pH 7.0). The resulting solution was stirred at 800 rpm for 12
using different isolation and processing methods may vary h (overnight) at room temperature to ensure complete dispersion and
widely in their molecular and physicochemical characteristics, dissolution. Pectin stock solution was then adjusted to pH 7 and
such as backbone length, electrical charge, hydrophobicity, equilibrated for 10 min before analysis.
surface activity, conformation, self-association, viscosity, and A stock emulsion was prepared by mixing 20% (w/w) corn oil and
binding capacity.17 In addition, the molecular and functional 80% (w/w) buffered emulsifier solution (5 mM phosphate buffer, pH
characteristics of pectins may be altered appreciably in response 7, containing 2.5% (w/w) Tween 80) together for 5 min using a
to changes in solution and environmental conditions such as biohomogenizer (speed 2, model MW140/2009-5, Biospec Products
pH, composition, ionic strength, and temperature.14,15,18,19 For Inc., ESGC, Switzerland). The resulting coarse emulsion was then
passed five times through a high-pressure homogenizer (microfluidizer
this reason, the study of the nature of the interactions that M-110L processor, Microfluidics Inc., Newton, MA, USA) operating at
pectin may have with the components of the gastrointestinal 11000 psi (75.8 MPa) to reduce the particle size further.
tract is important in understanding the effect of pectin on Pectin−emulsion mixtures were then prepared by mixing stock
digestive processes. An improved understanding of the origin emulsion (containing 20% (w/w) corn oil) with pectin stock solution
and nature of these interactions would lead to the design of (containing 4% (w/w) pectin) to obtain systems of various
functional foods with improved nutritional, physicochemical, compositions: 2% (w/w) corn oil and 0.2−3.6% (w/w) pectin. The
and sensory properties. pectin−emulsion mixtures were then stirred with a high-shear mixer
In this study, isothermal titration calorimetry (ITC) was used (Fisher Steadfast Stirrer, model SL-1200, Fisher Scientific, Pittsburgh,
to study the interactions between pectin and gastrointestinal PA, USA) at 1000 rpm and stored overnight at room temperature. The
pectin−emulsion mixtures were characterized to obtain the initial
components (bile salts, pancreatic lipase, CaCl2, and NaCl). phase, prior to subjection to the in vitro gastrointestinal model.
ITC measures the heat absorbed or evolved when one solution Simulated Gastrointestinal Tract Model. Each emulsion sample
is titrated into another solution.26−28 Previous studies have (initial phase) was passed through a simulated in vitro gastrointestinal
shown that ITC is an extremely valuable tool for studying tract that consisted of oral, gastric, and intestinal phases. This method
chemical interactions of polysaccharides with other molecules is based on a static model utilized in previous studies.31,32
because it provides valuable data concerning binding enthalpies, Simulated saliva fluid (SSF, pH 6.8) containing 3% (w/w) mucin
critical aggregation concentrations, and binding stoichiome- was prepared according to the composition shown in Table 1. Each
tries.29,30 In addition, electrophoresis, turbidity, and micro-
structural observations were used to provide additional Table 1. Chemical Composition of Simulated Saliva Fluid
information about the nature of the interactions between (SSF) Used To Simulate Oral Conditions
pectin and gastrointestinal components. The results of this concentrationa
study may aid the rational design of functional foods designed compound chemical formula (g/L)
to improve human health and wellness by controlling lipid sodium chloride NaCl 1.594
digestion within the gastrointestinal tract.


ammonium nitrate NH4NO3 0.328
potassium dihydrogen phosphate KH2PO4 0.636
MATERIALS AND METHODS potassium chloride KCl 0.202
Chemicals. Corn oil was purchased from a commercial food potassium citrate K3C6H5O7·H2O 0.308
supplier (Mazola, ACH Food Co. Inc., Memphis, TN, USA) and uric acid sodium salt C5H3N4O3Na 0.021
stored at 4 °C until use. The manufacturer reported that this oil
urea H2NCONH2 0.198
contained approximately 14, 29, and 57% (w/w) of saturated,
monounsaturated, and polyunsaturated fatty acids, respectively. lactic acid sodium salt C3H5O3Na 0.146
Commercial powdered high-methoxyl pectin (Genu Pectin (Citrus), porcine gastric mucin (type II) 30
a
USP/100) was kindly donated by CP Kelco (Lille Skensved, The SSF was prepared in double-distilled water and then pH 6.8 was
Denmark) and was used without further purification. The adjusted.
manufacturer reported that the powdered pectin ingredient contained
6.9% moisture, 89.0% galacturonic acid, and 8.6% methoxyl groups,
with a degree of esterification of approximately 62%. The average
molecular weight was reported to be 200 kDa. Lipase from porcine emulsion (initial phase) was mixed with SSF (ratio 1:1 w/w) to obtain
pancreas (type II, L3126, triacylglycerol hydrolase EC 3.1.1.3), bile a mixture containing 1% (w/w) corn oil and 0.1−1.8% (w/w) pectin.
extract (porcine, B8631), mucin from porcine stomach (type II, The simulated oral phase consisted of a conical flask containing an
M2378, bound sialic acid ≤ 1.2%), and pepsin A from porcine gastric emulsion−SSF mixture incubated at 37 °C with continuous shaking at
mucose (P7000, endopeptidase EC 3.4.23.1, activity ≥ 250 units/mg 100 rpm for 10 min in a temperature-controlled air incubator (Excella
solid) were purchased from Sigma-Aldrich Chemical Co. (St. Louis, E24 Incubator Shaker, New Brunswick Scientific, New Brunswick, NJ,
MO, USA). The supplier reported that the lipase activity was 100−400 USA). The mixture resulting from processing of the initial emulsions
units/mg protein (using olive oil) and 30−90 units/mg protein (using in the oral phase (bolus) was used in the gastric phase.

12621 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630


Journal of Agricultural and Food Chemistry Article

Simulated gastric fluid (SGF) was prepared by adding 2 g of NaCl, 7 lower amounts of the digestive components was prepared by diluting
mL of concentrated HCl (37% w/w), and 3.2 g of pepsin A (from the stock solution with buffer solution.
porcine gastric mucose, 250 units/mg) to a flask, then diluting the To establish the interaction of pectin with specific gastrointestinal
mixture with double-distilled water to a volume of 1 L, and finally components, individual solutions of CaCl2, NaCl, lipase, bile salts, and
adjusting the pH to 1.2 using 1 M HCl. Samples from the oral phase the mixture of all components were titrated into either buffer solution
(bolus) were mixed with SGF (ratio 1:1 w/w) so that the final mixture (blank) or pectin solution at 37 °C. Twenty-nine 100 μL aliquots of
contained 0.5% (w/w) corn oil and 0.05−0.9% (w/w) pectin. This individual solutions containing 2.7% (w/w) bile salts, 1.1% (w/w)
mixture was then adjusted to pH 2.5 using 1 M NaOH and incubated lipase, 55 mM CaCl2, 545 mM NaCl, or the mixture of all components
at 37 °C with continuous shaking at 100 rpm for 2 h. The mixture (100% relative concentration) were added sequentially into a test tube
resulting from processing of the emulsions in the gastric phase containing 14.5 mL of either buffer (5 mM phosphate buffer, pH 7) or
(chyme) was used in the intestinal phase. 0.9% (w/w) pectin solutions. Turbidity, ζ-potential, and micro-
Samples obtained from the gastric phase (20 mL) were incubated structure (observed by optical microscopy) of these solutions were
for 2 h at 37 °C in a simulated small intestine fluid (SIF) consisting of then characterized throughout the titration process.
2.5 mL of pancreatic lipase (24 mg/mL), 3.5 mL of bile extract Solution Characterization. Optical turbidity of the solutions was
solution (54 mg/mL), and 1.5 mL of salt solution containing 0.25 M determined by measuring their absorbance at 600 nm (Δτ = A600 nm)
CaCl2 and 3 M NaCl, to obtain a final composition in the reaction using a UV−visible spectrophotometer (Ultrospec 3000 pro
vessel of 0.36% (w/w) corn oil and 0.05−0.65% (w/w) pectin. The Pharmacia Biotech, Biochrom Ltd., Cambridge, UK) at 37 °C. The
free fatty acids (FFA) released were monitored by determining the samples were contained within 1 cm path length optical cells, and
amount of 0.1 M NaOH needed to maintain a constant pH 7.0 within buffer solution was used as a control. The change in turbidity was
the reaction vessel using an automatic titration unit (pH stat titrator, defined as Δτ = τpectin − τbuffer. Triplicate measurements of turbidity
835 Titrando, Metrohm USA, Inc.). All components were dissolved in were carried out on each sample.
5 mM phosphate buffer solution (pH 7) before use. Lipase addition The electrical charge (ζ-potential) of the solutions was determined
and initialization of the titration program were carried out only after using a particle microelectrophoresis instrument (Zetasizer NanoS-
the addition of all predissolved ingredients and balancing the pH to 7. eries, Malvern Instruments Ltd., Worcestershire, UK). Solutions were
The volume of 0.1 M NaOH added to the emulsion was recorded over injected into the measurement chamber and equilibrated for 120 s, and
time and used to calculate the FFAs generated by lipolysis. The then the ζ-potential was determined by measuring the direction and
amount of FFAs released was calculated using the equation velocity that the particles moved in the applied electric field. Each
individual ζ-potential measurement was calculated from the average of
FFA (% w/w) = 100 20 continuous readings made per sample. The ζ-potential was
recorded at each pH after 60 s of equilibrium.
⎛ VNaOH (L) × C NaOH (M) × MWlipid (g/mol) ⎞
The microstructure of the solutions was characterized by optical
× ⎜⎜ ⎟⎟
⎝ 2 × wlipid (g) ⎠ microscopy. An optical microscope (C1 Digital Eclipse, Niko, Tokyo,
Japan) with a 60× objective lens was used to capture images of the
(1) solutions. Solutions were gently stirred to form a homogeneous
where CNaOH is the concentration of the sodium hydroxide (0.1 M), mixture without introducing any air bubbles. A small aliquot of the
MWlipid is the average molecular weight of corn oil (872 g/mol), wlipid solutions (5 μL) was then transferred to a glass microscope slide and
is the initial weight of corn oil in the intestinal phase (0.1 g), and covered with a glass coverslip. The coverslip was fixed to the slide
VNaOH is the volume of NaOH (L) titrated into the reaction vessel to using nail polish to avoid evaporation. A small amount of immersion
neutralize the FFA released, assuming that all triacylglycerols are oil (type A, Nikon, Melville, NY, USA) was placed on the top of the
hydrolyzed in two molecules of FFA and one molecule of coverslip. All optical images were taken using a digital camera and then
monoacylglycerol. Titration blanks were performed by inactivating characterized using the instrument software (EZ CS1 version 3.8,
lipase in boiling water for 15 min prior to initialization of the titration Nikon).
program. Isothermal Titration Calorimetry Measurements. An ITC instru-
Creaming Stability Measurements of Initial Emulsions. Ten ment (Microcalorimeter VP-ITC, MicroCal Inc., Northampton, MA,
milliliters of emulsions (initial phase) was transferred into a test tube USA) was used to measure the enthalpy (ΔH) resulting from titration
(internal diameter = 15 mm, height = 125 mm), tightly sealed with a of individual solutions of CaCl2, NaCl, lipase, bile salts, and a mixture
plastic cap, and then stored at room temperature for 24 h, after which of all components (100% relative concentration) into either buffer
appreciable phase separation into an opaque layer at the top, a turbid solution (blank) or pectin solution. Twenty-nine 10 μL aliquots of
layer in the middle, and a transparent layer at the bottom was observed individual solutions of 2.7% (w/w) bile salts, 1.1% (w/w) lipase, 55
in some of the systems. We defined the serum layer to be the sum of mM calcium chloride, 545 mM sodium chloride, or a mixture of all
the turbid and transparent layers. The total height of the emulsion components (100% relative concentration) were injected sequentially
(HE) and the height of the serum layer (HS) were measured using a into a 1450 μL titration cell initially containing either buffer (5 mM
laser vertical profiling system (Turbiscan Classic MA2000, For- phosphate buffer, pH 7) or 0.9% (w/w) pectin solutions. The range of
mulaction, Wynnewood, PA, USA). The extent of creaming was then concentrations of pectin and gastrointestinal components used was
characterized by the creaming index (CI), defined as CI = 100 × (HS/ selected to approximately mimic the concentrations in the small
HE). The creaming index provided indirect information about droplet intestinal phase of the gastrointestinal tract. Each injection lasted 24 s,
aggregation, because an increase in particle size (e.g., due to and there was an interval of 240 s between successive injections. The
flocculation) leads to faster creaming (provided the droplet temperature of the solution in the titration cell was 37 °C, and the
concentration is not too high). solution was stirred at a speed of 315 rpm throughout the experiments.
Interaction of Pectin with Gastrointestinal Components. Triplicate measurements of enthalpy profiles were carried out on each
Solution Preparation. Stock solutions of 0.9% (w/w) pectin, 2.7% sample, and they were reproducible to better than 5%.
(w/w) bile salts, and 1.1% (w/w) lipase were individually prepared by Data Analysis. All measurements were performed at least three
dispersing them in 5 mM phosphate buffer solution (pH 7) at room times using freshly prepared samples. Averages and standard deviations
temperature for 12 h (overnight), to ensure their complete dispersion were calculated from these triplet measurements.


and dissolution. Salt solutions consisting of 55 mM CaCl2 or 545 mM
NaCl were prepared freshly before each experiment. A stock solution RESULTS AND DISCUSSION
containing a mixture of all the above components at the same initial
concentrations as in their individual solutions was also prepared. This Properties of Initial Emulsion−Pectin Mixtures.
stock solution was referred to as having a relative concentration of Initially, the properties of the initial emulsions used to prepare
“100%” of the digestive components. A series of mixtures containing the emulsion−pectin mixtures were characterized (5 mM
12622 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630
Journal of Agricultural and Food Chemistry Article

phosphate buffer, pH 7). The initial emulsions had a relatively increased from 0 to 0.2% (w/w). However, emulsions
small mean particle diameter immediately after preparation (d32 containing higher levels of pectin clearly separated into a
= 200 ± 10 nm), which did not change during the course of the creamed layer and a serum layer, indicating that the droplets
experiments.33 This can be attributed to the relatively strong moved upward due to gravity. The origin of this effect is
steric repulsion between the nonionic surfactant-coated lipid depletion flocculation induced by the presence of nonadsorbed
droplets resulting from the polyoxyethylene headgroups on the pectin molecules within the aqueous phase.24,38 There was an
Tween 80 molecules.34 The lipid droplets initially had a small appreciable increase in mean particle diameter in the samples in
negative charge (ζ = −5.8 ± 0.2 mV), even though they were which depletion flocculation occurred, for example, d32 = 350
stabilized by a nonionic surfactant, which can be attributed to nm in the sample initially containing 2.4% (w/w) pectin. This
anionic impurities in the oil or surfactant ingredients (such as effect can be attributed to the fact that the droplets were
free fatty acids) or to preferential adsorption of hydroxyl ions pushed together in the flocs due to the depletion attraction,
from water (OH−).35,36 The droplets in the initial emulsions which promoted coalescence. Pectin molecules should not be
containing no pectin were stable to creaming throughout the attracted to the surfaces of the nonionic surfactant-coated lipid
experimental time frame (Figure 1), which is due to the fact droplets because of electrostatic and steric repulsion effects (the
that the droplets were small and did not associate with each carboxyl groups of pectin are completely dissociated at pH 7
other.37 because its pKa is ∼3.5). There is therefore a narrow region
around each lipid droplet (approximately equal to the radius of
hydration of the pectin molecules) from which the pectin
molecules are excluded.39 Consequently, there is a pectin and
water concentration gradient between this exclusion zone and
the surrounding bulk aqueous phase, which leads to an osmotic
pressure. This osmotic pressure generates an attraction between
the lipid droplets (depletion force) because when they come
into close contact, the volume of the thermodynamically
unfavorable exclusion zone is reduced.25 At low pectin
concentrations, the osmotic attraction is not large enough to
overcome the various repulsive forces in the system (e.g., steric
and electrostatic), and so the emulsion remains stable to
droplet aggregation. However, once a critical pectin concen-
tration is exceeded (∼0.5% w/w), the attractive forces exceed
the repulsive forces and droplet flocculation occurs.24 It should
be noted that the main thermodynamic driving force for
depletion flocculation is entropy.40,41 The reason that the
Figure 1. Influence of pectin concentration on creaming stability of 2% thickness of the creamed layer increases (CI decreases) as the
(w/w) corn oil-in-water emulsions stabilized with 0.2% (w/w) Tween pectin concentration increases above the critical flocculation
80 before the digestion process. concentration (Figure 1) is that the depletion attraction is
higher and so the droplets are held more strongly into a particle
Emulsions containing relatively low levels of added pectin gel network.42 In summary, these measurements showed that
(≤0.4% w/w) were still observed to be stable to gravitational the degree of droplet flocculation in the initial emulsion−pectin
separation; that is, they maintained a uniform white appearance mixtures depended on the amount of pectin present in the
(Figure 1). They also had good stability to droplet aggregation system.
with the mean particle diameter (d32) increasing only from Influence of Pectin on Lipid Digestion. In a recent
around 200 to 210 nm when the pectin concentration was study, we examined the influence of pectin addition on the

Figure 2. Influence of pectin concentration on free fatty acids (FFA) released during lipid digestion process (pH stat): (a) kinetics; (b) FFA released
after 2 h. Pectin concentrations refer to those present in the intestinal phase.

12623 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630


Journal of Agricultural and Food Chemistry Article

gastrointestinal fate of emulisified lipids.33 This study showed


that pectin influenced the aggregation state of the lipid droplets
in different regions of the simulated gastrointestinal tract
(mouth, stomach, and small intestine). It also showed that
pectin addition reduced the rate and extent of lipase-catalyzed
lipid digestion in the small intestine phase. In the current study,
the same pH-stat method was used to measure the influence of
pectin on lipid digestion so that we could directly compare
these results with the studies of pectin interactions with various
gastrointestinal components using the same constituents.
In general, the amount of FFAs produced increased rapidly
during the first few minutes of digestion and then increased
more slowly at longer times, until a relatively constant value Figure 3. Influence of pectin on the microstructure (observed by
was reached after 25 min of digestion (Figure 2a). The final optical microscopy) of emulsions before (initial phase) and after
(intestinal phase) of lipid digestion. The initial emulsions contained
amount of FFAs produced at the end of the 2 h digestion
2% (w/w) corn oil, 2% (w/w) Tween 80, and 0 or 2.4% (w/w) pectin.
period decreased with increasing pectin concentration (Figure The scale bar corresponds to 20 μm.
2b), which is in good agreement with our earlier study.33
The ability of pectin to decrease the rate and extent of lipid
digestion in corn oil-in-water emulsions could be due to several
physicochemical and physiological mechanisms. Pectin may be observed by optical microscopy, but confocal microscopy
have interacted with calcium ions and formed a highly viscous images have shown that they are evenly distributed throughout
solution or gel network that impeded the diffusion of GIT the system.33 After digestion, there were clearly some large
components (such as lipase or bile salts), thereby retarding the aggregates in the control samples, which may have been lipid
lipid digestion process.43−45 Pectin−calcium interactions may digestion products or remnants from the various GIT
also have inhibited lipid digestion due to the important role components (such as calcium, sodium, bile salts, and lipase).
that calcium ions play in removing FFAs from lipid droplet In the presence of pectin, the lipid droplets were trapped within
surfaces.46 Calcium ions normally form insoluble soaps with large clusters prior to digestion, which can be attributed to
long-chain FFAs that help remove the FFAs from the lipid depletion flocculation induced by nonadsorbed pectin mole-
droplet surfaces and thereby allow the lipase to keep working.47 cules (Figure 3).25 After digestion, there were a relatively high
If the calcium ions are tightly bound to pectin molecules, then number of large aggregates (undigested lipid droplets)
the FFAs may accumulate at the lipid droplet surfaces, thereby remaining in the samples containing pectin, which could have
inhibiting further digestion. Pectin may also have bound to bile been due to interactions of the pectin molecules with various
salts or phospholipids and so prevented them from adsorbing components in the simulated GIT fluids.
to lipid droplet surfaces or from solubilizing lipid digestion Turbidity (Δτ), ζ-potential, and enthalpy change (ΔH) were
products.10,18,44 The solubilization of long-chain FFAs in mixed measured when increasing amounts of mixed GIT fluids (0 to
micelles is another important means of removing them from 100% of mixture) were titrated into a pectin (0.9% w/w)
the lipid droplet surfaces and thereby allowing lipase to keep solution at pH 7 (Figure 4). The 100% GIT fluids contained
functioning.34 The adsorption of bile salts and phospholipids to the levels of the NaCl, CaCl2, bile salts, and lipase in the final
lipid droplet surfaces often facilitates the subsequent adsorption simulated small intestinal fluids. There was a progressive
of lipase molecules. Pectin may also be able to alter lipid increase in the turbidity when the amount of mixed GIT fluids
digestion by interacting directly with lipase molecules.21,48 titrated into the pectin solution increased (Figure 4a). The ζ-
Finally, the presence of pectin in the system may alter the potential of the mixed system moved from highly negative
aggregation state of the lipid droplets through either depletion (−19 mV) to less negative (−2 mV) as increasing amounts of
or bridging flocculation.49 The digestion of highly flocculated GIT fluids were added (Figure 4a), which may have occurred
lipid droplets is often less than that of nonflocculated ones due to electrostatic screening effects induced by the presence of
because the lipase molecules have to diffuse through the flocs salts or due to binding of cations in the GIT fluids to anionic
before they can reach the lipid droplets in their interiors.23,50 pectin molecules.51 There was also an appreciable difference in
Clearly, the potential influence of pectin on the lipid digestion the enthalpy profiles of samples with either buffer or pectin
process within the GIT is complex because pectin is able to solutions (Figure 4b). In the absence of pectin, the enthalpy
interact with many constituents within the GIT fluids. change was relatively low and may be attributed to heat of
The purpose of the current study was to provide further dilution effects associated with the various molecules and/or
insights into the potential origin of these effects by studying the particles moving further apart when the GIT fluids were
interactions between pectin and specific gastrointestinal injected into the reaction cell.22,27,30 In the presence of pectin,
components. there was a large exothermic enthalpy change when the first few
Interactions of Pectin with Mixed Gastrointestinal aliquots of the mixed GIT fluids were injected into the reaction
Components. In this series of experiments, we used ITC, cell containing the pectin solution. The enthalpy change
electrophoresis, turbidity, and microstructure measurements to became progressively less exothermic when 0−20% of the GIT
examine the interactions of pectin with a solution containing a fluids was added, until it eventually reached a relatively constant
mixture of all the major GIT components: sodium, calcium, bile endothermic value at higher GIT levels (∼40%). These results
salts, and lipase. The microstructure of the emulsion−pectin suggest that the pectin molecules interacted with some of the
mixtures changed appreciably after exposure to the mixed components in the GIT fluids, possibly through electrostatic
gastrointestinal components (Figure 3). In the absence of interactions, and formed aggregates that were large enough to
pectin (control), the individual lipid droplets were too small to scatter light strongly.
12624 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630
Journal of Agricultural and Food Chemistry Article

Figure 4. Influence of relative concentration (0−100%) of simulated gastrointestinal fluids on (a) optical turbidity (Δτ = A600 nm) and ζ-potential
and (b) interaction enthalpy (ΔH) of solutions containing either buffer solution or an initial concentration of 0.9% (w/w) pectin. The initial
simulated gastrointestinal fluids (100%) contained 0.45% (w/w) bile salts, 0.18% (w/w) lipase, 9 mM CaCl2, and 91 mM NaCl. Optical turbidity
(Δτ) was defined as Δτ = τpectin − τbuffer.

Figure 5. Influence of the concentrations of NaCl, CaCl2, bile salts, and lipase on optical turbidity (Δτ = A600 nm) and ζ-potential of solutions
containing an initial concentration of 0.9% (w/w) pectin.

Interactions of Pectin with Specific Gastrointestinal In addition, the turbidities and ζ-potentials reached at the end
Components. In this series of experiments, ITC, electro- of these titrations were reported so that different samples could
phoresis, turbidity, and microstructure measurements were be compared more easily (Figure 6).
used to characterize the interactions of pectin with specific GIT NaCl. Addition of increasing amounts of NaCl to the pectin
components (NaCl, CaCl2, bile salts, or lipase). Measurements solutions caused an appreciable decrease in the magnitude of
were made as increasing amounts of GIT components were the ζ-potential of the pectin solutions (Figure 5, NaCl), which
individually added to a pectin solution (0.9% w/w) (Figure 5). can be attributed to electrostatic screening effects, that is,
12625 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630
Journal of Agricultural and Food Chemistry Article

Figure 6. Influence of 91 mM NaCl, 9 mM CaCl2, 0.45% (w/w) bile salts (BS), 0.18% (w/w) lipase (L), and a mixture of all components (M, 100%
relative concentrations) on optical turbidity (Δτ = A600 nm) (a) and ζ-potential (b) of solutions containing an initial concentration of 0.9% (w/w)
pectin (P).

accumulation of sodium ions around the negative groups on the


pectin molecules.51 There was no evidence of an increase in
aggregate formation within the microstructure images (Figure
8, NaCl), and there was little change in the general appearance
(Figure 7, NaCl) and optical turbidity (Figure 5, NaCl) of these

Figure 8. Influence of 91 mM NaCl, 9 mM CaCl2, 0.45% (w/w) bile


salts (BS), 0.18% (w/w) lipase (L), and a mixture of the previous
components (100% relative concentrations) on the microstructure
(observed by optical microscopy) of solutions prepared on either
buffer solution or 0.9% (w/w) pectin. The scale bar corresponds to 20
μm.

and the NaCl (Figure 9, NaCl), because the enthalpy versus salt
concentration profiles were fairly similar. Overall, these results
Figure 7. Influence of the concentrations of NaCl, CaCl2, bile salts, suggest that pectin did not have a major interaction with the
lipase, and a mixture of all components (% relative concentrations) on NaCl in the GIT fluids.
the turbidity of solutions containing an initial concentration of 0.9% CaCl2. Addition of CaCl2 to the pectin solutions had a more
(w/w) pectin. pronounced influence on their physicochemical properties than
NaCl addition. Addition of increasing amounts of CaCl2 caused
the ζ-potential of the pectin solutions to go from around −19
samples upon NaCl addition. We did observe a slight increase mV to around −6 mV (Figure 5, CaCl2), which can be
in the turbidity at the highest NaCl levels used, which suggested attributed to electrostatic screening and ion binding effects.51
that the NaCl may have caused a slight amount of pectin Multivalent counterions are more effective at screening
aggregation, presumably through screening of the electrostatic electrostatic interactions than monovalent ones, and they
interactions.52 The ITC measurements indicated that there was have a greater tendency to bind strongly to oppositely charged
not a strong interaction (ΔH = −102 kcal) between the pectin groups.53 The change in the ζ-potential with calcium addition
12626 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630
Journal of Agricultural and Food Chemistry Article

Figure 9. Influence of the concentrations of NaCl, CaCl2, bile salts, and lipase on interaction enthalpy (ΔH) of solutions containing either buffer
solution or an initial concentration of 0.9% (w/w) pectin.

was fairly similar to that observed with the addition of the mentioned earlier, the binding of calcium ions to pectin may be
mixed GIT solution (Figure 6b), which suggests that calcium important for a number of reasons. First, pectin−calcium
ions play an important role in determining the overall electrical microgels may trap lipid droplets inside, thereby restricting the
interactions. The ITC measurements also indicated that there access of lipase to the lipid substrates.23 Second, the formation
was a strong interaction (ΔH = −241 kcal) between the of these electrostatic complexes may prevent the calcium ions
calcium ions and pectin (Figure 9, CaCl2). When calcium ions from forming insoluble soaps with long-chain FFAs at the lipid
were injected into buffer solution, there was an exothermic droplet surfaces, thereby preventing the lipase from functioning
enthalpy change that decreased in magnitude with increasing properly.47 In practice, there will be other types of salts in the
CaCl2 concentration, which can be attributed to heat of dilution gastrointestinal fluids that may alter the interactions of calcium
effects.22 However, when calcium ions were injected into pectin with pectin through electrostatic screening or binding effects. It
solution, there was initially a strong exothermic reaction (from should be noted that there may have been some sodium or
0 to 4 mM), then a relatively strong endothermic reaction calcium present in the ingredients used in this study (such as
(from 4 to 6 mM), followed by an enthalpy change close to pectin), and these ions could also contribute to the observed
zero after 6 mM CaCl2 (Figure 9, CaCl2). The shape of the electrostatic interactions.
ITC curve suggests that there may have been several events Bile Salts. The addition of bile salts to the pectin solutions
occurring sequentially, such as calcium binding, pectin had a limited influence on their physicochemical properties.
conformational changes, and/or pectin aggregation. Never- Addition of increasing amounts of bile salts caused only a
theless, it is not possible to determine the molecular origin of relatively small decrease in the negative charge on the pectin
these events from ITC measurements alone because they molecules (Figure 5, bile salts), which can probably be
provide only the overall enthalpy change. attributed to electrostatic screening effects.7,9 There was an
The addition of calcium ions to the pectin solutions caused a appreciable increase in the optical turbidity of the pectin
large increase in their optical turbidity (Figures 5 (CaCl2) and solutions when bile salts were added (Figures 5 (bile salts) and
6a), which can be attributed to the appearance of large 6a), and the solutions appeared visibly cloudier (Figure 7, bile
aggregates, as observed within the optical microscopy images salts). These effects can be attributed to the formation of large
(Figures 7 and 8, CaCl2). We postulate that these aggregates aggregates within the bile salts−pectin solutions as observed by
were microgels formed by pectin in the presence of optical microscopy (Figure 8, bile salts). The ITC measure-
calcium.43,45 The calcium ions may have cross-linked the ments also indicated that the bile salts interacted with the
anionic pectin molecules through electrostatic bridging.13 As pectin molecules. In the absence of pectin, there was a relatively
12627 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630
Journal of Agricultural and Food Chemistry Article

large endothermic enthalpy change (ΔH = 204 kcal), which associated with the interaction (endothermic or exothermic).
progressively decreased with increasing bile salt concentration For more complex systems, the interpretation of ITC data is
(Figure 9, bile salts). This effect may have been due to the more difficult because many different molecular and
breakdown of bile salt micelles within the reaction chamber, physicochemical events contribute to the overall enthalpy
resulting in the exposure of a greater number of nonpolar changes, such as binding, conformation changes, and
groups to water.54 In the presence of pectin, there was a large aggregation. Consequently, it is not possible to establish the
endothermic peak (from 0.05 to 0.15% w/w) followed by a precise nature of the different contributions to the measured
value that was close to zero at higher bile salt concentrations ITC signal. Nevertheless, ITC measurements do still provide
(after 0.2% (w/w), Figure 9, bile salts). The large difference valuable information, because they indicate that one or more
between the two curves suggests that there was an interaction types of interaction are taking place between two molecules,
between pectin and bile salts; for example, bile salt micelles may and they show the range of concentrations over which these
have bound to pectin molecules and/or promoted their intearctions occur. Typically, ITC measurements should be
aggregation.55 combined with other techniques to provide a more detailed
One might not expect a strong interaction between bile salts understanding of the nature of the interactions involved, such
and pectin molecules because they are both usually negatively as chromatography, scattering, spectrometry, or spectroscopy
charged at neutral pH. However, bile salts may be able to methods.
interact with pectin molecules through hydrophobic inter- The purpose of this study was to investigate the interaction
actions.56 In general, the hydrophobic interactions of pectin of pectin molecules with some of the major constituents within
with bile salts are known to increase with the degree of gastrointestinal fluids, that is, NaCl, CaCl2, bile salts, and lipase.
esterification of pectin molecules (by increasing its hydro- Our measurements have shown that a number of these
phobicity) and the molecular weight (viscosity) of the pectin components interact strongly with pectin, which may have
molecules.20 These experiments clearly show that bile salts can important implications for understanding the influence of
interact with pectin molecules, which may have important pectin on lipid digestion. In particular, calcium ions and bile
implications for the effects of pectin on lipid digestion observed salts appear to promote the formation of pectin microgels that
in Figure 2. If the bile salts form microgel particles with pectin, can lead to flocculation of lipid droplets in the gastrointestinal
then they may not be available to absorb to the surfaces of the tract, decreasing the ability of emulsified lipids to be accessed
lipid droplets or to solubilize lipid digestion products, as by gastrointestinal lipases and thereby inhibiting digestion. If
discussed earlier. In addition, if any lipid droplets are trapped appreciable amounts of calcium and bile salts are trapped within
inside the microgels, then the rate of digestion may be these microgels, then they will not be able to remove long-chain
decreased because the lipase molecules cannot easily reach the fatty acids from lipid droplet surfaces, which would inhibit lipid
lipid droplet surfaces.9 digestion. In addition, if lipid droplets are trapped within pectin
Lipase. Finally, we examined the influence of lipase addition microgels, then it may be more difficult for lipase molecules to
to the pectin solutions on their overall physicochemical access the surfaces of the lipid substrate, again delaying lipid
properties. The addition of increasing amounts of lipase caused digestion. These results have important implications for the
a slight change in the ζ-potential of the pectin (Figure 5, rational design of dietary fiber-based functional foods that may
modulate lipid digestion within the gastrointestinal tract.


lipase), which may have been due to some binding of lipase
molecules to the pectin molecules. In addition, there was an
appreciable difference between the ITC curves of the control AUTHOR INFORMATION
and the system containing pectin (Figure 9, lipase), which also Corresponding Author
suggests that some form of interaction occurred (ΔH = −129 *(D.J.M.) E-mail: mcclements@foodsci.umass.edu. Phone:
kcal).21 In the absence of pectin, there was a relatively large (413) 545-1019. Fax: (413) 545-1262.
exothermic change observed when the lipase solution was Funding
titrated into the buffer solution. This effect can probably be We are grateful to COLCIENCIAS and Universidad Nacional
attributed to the heat of dilution associated with salts in the de Colombia for providing a fellowship to M.E.-R. supporting
lipase solution. In the presence of pectin, there was a large this work. This material is partly based upon work supported by
exothermic change observed at low lipase levels (0−0.1 w/w the U.S. Department of Agriculture, NRI Grants (2011-03539,
%), followed by a much smaller exothermic change at higher 2013-03795, 2011-67021, and 2014-67021).
lipase levels. The large difference between the curves in the Notes
presence and absence of pectin suggests that there was a strong The authors declare no competing financial interest.


interaction between the lipase and the pectin, although again
the molecular origin of this effect cannot be established from ABBREVIATIONS USED
the ITC measurements. There was a slight increase in the FFA, free fatty acid; GalA, D-galacturonic acid; GIT, gastro-
optical turbidity of the lipase−pectin solutions when high levels intestinal tract; ITC, isothermal titration calorimetry; SSF,
of lipase were added (Figure 5, lipase), but there was little simulated saliva fluid; SGF, simulated gastric fluid; SIF,
change in their overall visual appearance (Figure 7, lipase) or simulated intestinal fluid


their microstructure (Figure 8, lipase). These results suggest
either that any interactions between the lipase and pectin did REFERENCES
not lead to extensive aggregation or that the lipase−pectin
(1) Bray, G. A.; Popkin, B. M. Dietary fat intake does affect obesity!
complexes formed were soluble in buffer solution. Am. J. Clin. Nutr. 1998, 68, 1157−1173.
For simple systems involving the binding of a ligand to (2) Howarth, N. C.; Saltzman, E.; Roberts, S. B. Dietary fiber and
another molecule, ITC is a powerful tool for determining weight regulation. Nutr. Rev. 2001, 59, 129−139.
thermodynamic properties, such as the number and strength of (3) Beysseriat, M.; Decker, E. A.; McClements, D. J. Preliminary
binding sites and the nature of the enthalpy change (ΔH) study of the influence of dietary fiber on the properties of oil-in-water

12628 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630


Journal of Agricultural and Food Chemistry Article

emulsions passing through an in vitro human digestion model. Food (26) Leavitt, S.; Freire, E. Direct measurement of protein binding
Hydrocolloids 2006, 20, 800−809. energetics by isothermal titration calorimetry. Curr. Opin. Struct. Biol.
(4) Mun, S.; Decker, E.; Park, Y.; Weiss, J.; McClements, D. J. 2001, 11, 560−566.
Influence of interfacial composition on in vitro digestibility of (27) Doyle, M. L. Characterization of binding interactions by
emulsified lipids: potential mechanism for chitosan’s ability to inhibit isothermal titration calorimetry. Curr. Opin. Biotechnol. 1997, 8, 31−
fat digestion. Food Biophys. 2006, 1, 21−29. 35.
(5) McClements, D. J.; Li, Y. Structured emulsion-based delivery (28) Freire, E.; Mayorga, O. L.; Straume, M. Isothermal titration
systems: controlling the digestion and release of lipophilic food calorimetry. Anal. Chem. 1990, 62, 950A−959A.
components. Adv. Colloid Interface Sci. 2010, 159, 213−228. (29) Wangsakan, A.; Chinachoti, P.; McClements, D. J. Effect of
(6) Gunness, P.; Gidley, M. J. Mechanisms underlying the surfactant type on surfactant−maltodextrin interactions: isothermal
cholesterol-lowering properties of soluble dietary fibre polysaccharides. titration calorimetry, surface tensiometry, and ultrasonic velocimetry
Food Funct. 2010, 1, 149−155. study. Langmuir 2004, 20, 3913−3919.
(7) Hur, S. J.; Lim, B. O.; Decker, E. A.; McClements, D. J. In vitro (30) Chang, Y.; McLandsborough, L.; McClements, D. J. Interactions
human digestion models for food applications. Food Chem. 2011, 125, of a cationic antimicrobial (ε-polylysine) with an anionic biopolymer
(pectin): an isothermal titration calorimetry, microelectrophoresis, and
1−12.
(8) Li, Y.; McClements, D. J. Modulating lipid droplet intestinal turbidity study. J. Agric. Food Chem. 2011, 59, 5579−5588.
(31) Hur, S. J.; Decker, E. A.; McClements, D. J. Influence of initial
lipolysis by electrostatic complexation with anionic polysaccharides:
emulsifier type on microstructural changes occurring in emulsified
Influence of cosurfactants. Food Hydrocolloids 2014, 35, 367−374.
lipids during in vitro digestion. Food Chem. 2009, 114, 253−262.
(9) McClements, D. J.; Li, Y. Review of in vitro digestion models for
(32) Salvia-Trujillo, L.; Qian, C.; Martin-Belloso, O.; McClements,
rapid screening of emulsion-based systems. Food Funct. 2010, 1, 32− D. J. Influence of particle size on lipid digestion and beta-carotene
59. bioaccessibility in emulsions and nanoemulsions. Food Chem. 2013,
(10) Anderson, J. W.; Baird, P.; Davis, R. H., Jr; Ferreri, S.; Knudtson, 141, 1472−1480.
M.; Koraym, A.; Waters, V.; Williams, C. L. Health benefits of dietary (33) Espinal-Ruiz, M.; Parada-Alfonso, F.; Restrepo-Sánchez, L.-P.;
fiber. Nutr. Rev. 2009, 67, 188−205. Narváez-Cuenca, C.-E.; McClements, D. J. Impact of dietary fibers
(11) Kritchevsky, D. Dietary fiber. Annu. Rev. Nutr. 1988, 8, 301− [methyl cellulose, chitosan, and pectin] on digestion of lipids under
328. simulated gastrointestinal conditions. Food Funct. 2014, 5, 3083−3095.
(12) Theuwissen, E.; Mensink, R. P. Water-soluble dietary fibers and (34) Almgren, M. Mixed micelles and other structures in the
cardiovascular disease. Physiol. Behav. 2008, 94, 285−292. solubilization of bilayer lipid membranes by surfactants. Biochim.
(13) Thakur, B. R.; Singh, R. K.; Handa, A. K.; Rao, M. A. Chemistry Biophys. Acta−Biomembranes 2000, 1508, 146−163.
and uses of pectin − a review. Crit. Rev. Food Sci. Nutr. 1997, 37, 47− (35) Thiam, A. R.; Farese, R. V., Jr.; Walther, T. C. The biophysics
73. and cell biology of lipid droplets. Nat. Rev. Mol. Cell Biol. 2013, 14,
(14) Mohnen, D. Pectin structure and biosynthesis. Curr. Opin. Plant 775−786.
Biol. 2008, 11, 266−277. (36) Singh, H.; Ye, A.; Horne, D. Structuring food emulsions in the
(15) Ridley, B. L.; O’Neill, M. A.; Mohnen, D. Pectins: structure, gastrointestinal tract to modify lipid digestion. Prog. Lipid Res. 2009,
biosynthesis, and oligogalacturonide-related signaling. Phytochemistry 48, 92−100.
2001, 57, 929−967. (37) Weiss, J.; Takhistov, P.; McClements, D. J. Functional materials
(16) Caffall, K. H.; Mohnen, D. The structure, function, and in food nanotechnology. J. Food Sci. 2006, 71, R107−R116.
biosynthesis of plant cell wall pectic polysaccharides. Carbohydr. Res. (38) Dickinson, E.; Semenova, M. G.; Antipova, A. S.; Pelan, E. G.
2009, 344, 1879−1900. Effect of high-methoxy pectin on properties of casein-stabilized
(17) Yapo, B. M. Pectic substances: from simple pectic emulsions. Food Hydrocolloids 1998, 12, 425−432.
polysaccharides to complex pectins − a new hypothetical model. (39) Jenkins, P.; Snowden, M. Depletion flocculation in colloidal
Carbohydr. Polym. 2011, 86, 373−385. dispersions. Adv. Colloid Interface Sci. 1996, 68, 57−96.
(18) Maxwell, E. G.; Belshaw, N. J.; Waldron, K. W.; Morris, V. J. (40) Reiffers-Magnani, C. K.; Cuq, J. L.; Watzke, H. J. Depletion
Pectin − an emerging new bioactive food polysaccharide. Trends Food flocculation and thermodynamic incompatibility in whey protein
Sci. Technol. 2012, 24, 64−73. stabilised O/W emulsions. Food Hydrocolloids 2000, 14, 521−530.
(19) Munarin, F.; Tanzi, M. C.; Petrini, P. Advances in biomedical (41) Méndez-Alcaraz, J. M.; Klein, R. Depletion forces in colloidal
applications of pectin gels. Int. J. Biol. Macromol. 2012, 51, 681−689. mixtures. Phys. Rev. E 2000, 61, 4095−4099.
(20) Pfeffer, P. E.; Doner, L. W.; Hoagland, P. D.; McDonald, G. G. (42) Dickinson, E. On flocculation and gelation in concentrated
Molecular interactions with dietary fiber components. Investigation of particulate systems containing added polymer. J. Chem. Soc., Faraday
Trans. 1995, 91, 4413−4417.
the possible association of pectin and bile acids. J. Agric. Food Chem.
(43) Willats, W. G. T.; Knox, J. P.; Mikkelsen, J. D. Pectin: new
1981, 29, 455−461.
insights into an old polymer are starting to gel. Trends Food Sci.
(21) Espinal-Ruiz, M.; Parada-Alfonso, F.; Restrepo-Sánchez, L.-P.;
Technol. 2006, 17, 97−104.
Narváez-Cuenca, C.-E. Inhibition of digestive enzyme activities by (44) Leroux, J.; Langendorff, V.; Schick, G.; Vaishnav, V.; Mazoyer, J.
pectic polysaccharides in model solutions. Bioact. Carbohydr. Dietary Emulsion stabilizing properties of pectin. Food Hydrocolloids 2003, 17,
Fibre 2014, 4, 27−38. 455−462.
(22) McClements, D. J. Isothermal titration calorimetry study of (45) Braccini, I.; Pérez, S. Molecular basis of Ca2+-induced gelation in
pectin−ionic surfactant interactions. J. Agric. Food Chem. 2000, 48, alginates and pectins: the egg-box model revisited. Biomacromolecules
5604−5611. 2001, 2, 1089−1096.
(23) Simo, O. K.; Mao, Y.; Tokle, T.; Decker, E. A.; McClements, D. (46) Devraj, R.; Williams, H. D.; Warren, D. B.; Mullertz, A.; Porter,
J. Novel strategies for fabricating reduced fat foods: heteroaggregation C. J. H.; Pouton, C. W. In vitro digestion testing of lipid-based delivery
of lipid droplets with polysaccharides. Food Res. Int. 2012, 48, 337− systems: calcium ions combine with fatty acids liberated from
345. triglyceride rich lipid solutions to form soaps and reduce the
(24) Jenkins, P.; Snowden, M. Depletion flocculation in colloidal solubilization capacity of colloidal digestion products. Int. J. Pharm.
dispersions. Adv. Colloid Interface Sci. 1996, 68, 57−96. 2013, 441, 323−333.
(25) McClements, D. J. Comments on viscosity enhancement and (47) Ye, A.; Cui, J.; Zhu, X.; Singh, H. Effect of calcium on the
depletion flocculation by polysaccharides. Food Hydrocolloids 2000, 14, kinetics of free fatty acid release during in vitro lipid digestion in
173−177. model emulsions. Food Chem. 2013, 139, 681−688.

12629 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630


Journal of Agricultural and Food Chemistry Article

(48) de Roos, A.; Walstra, P. Loss of enzyme activity due to


adsorption onto emulsion droplets. Colloids Surf., B 1996, 6, 201−208.
(49) Biggs, S.; Habgood, M.; Jameson, G. J.; Yan, Y.-d. Aggregate
structures formed via a bridging flocculation mechanism. Chem. Eng. J.
2000, 80, 13−22.
(50) Li, Y.; Hu, M.; McClements, D. J. Factors affecting lipase
digestibility of emulsified lipids using an in vitro digestion model:
proposal for a standardised pH-stat method. Food Chem. 2011, 126,
498−505.
(51) Israelachvili, J. Intermolecular and Surface Forces, 3rd ed.;
Academic Press: London, UK, 2011.
(52) Furusawa, K.; Ueda, M.; Nashima, T. Bridging and depletion
flocculation of synthetic latices induced by polyelectrolytes. Colloids
Surf., A: Physicochem. Eng. Aspects 1999, 153, 575−581.
(53) Winzor, D. J.; Carrington, L. E.; Deszczynski, M.; Harding, S. E.
Extent of charge screening in aqueous polysaccharide solutions.
Biomacromolecules 2004, 5, 2456−2460.
(54) Thongngam, M.; McClements, D. J. Isothermal titration
calorimetry study of the interactions between chitosan and a bile
salt (sodium taurocholate). Food Hydrocolloids 2005, 19, 813−819.
(55) Wangsakan, A.; Chinachoti, P.; McClements, D. J. Isothermal
titration calorimetry study of the influence of temperature, pH and salt
on maltodextrin−anionic surfactant interactions. Food Hydrocolloids
2006, 20, 461−467.
(56) Dongowski, G. Effect of pH on the in vitro interactions between
bile acids and pectin. Z. Lebensm. Unters. Forsch. 1997, 205, 185−192.

12630 dx.doi.org/10.1021/jf504829h | J. Agric. Food Chem. 2014, 62, 12620−12630

You might also like